Next Article in Journal
Centromere Repeats: Hidden Gems of the Genome
Next Article in Special Issue
Exceptional Longevity and Polygenic Risk for Cardiovascular Health
Previous Article in Journal
DNA Replication Timing Enters the Single-Cell Era
Previous Article in Special Issue
Lifespans of Twins: Does Zygosity Matter?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Genetic Variability of APOE in Different Human Populations and Its Implications for Longevity

1
Laboratory of Molecular Anthropology & Centre for Genome Biology, Department of Biological, Geological and Environmental Sciences, University of Bologna, 40126 Bologna, Italy
2
Department of Experimental, Diagnostic, and Specialty Medicine (DIMES), University of Bologna, 40126 Bologna, Italy
3
Department of Experimental and Clinical Biomedical Sciences “Mario Serio”, University of Florence, 50134 Florence, Italy
4
IRCCS Istituto delle Scienze Neurologiche di Bologna, 40139 Bologna, Italy
5
Department of Cultural Heritage (DBC), University of Bologna, Ravenna Campus, 48121 Ravenna, Italy
6
School of Anthropology and Museum Ethnography, University of Oxford, OX2 6PE Oxford, UK
*
Authors to whom correspondence should be addressed.
Equally contributing authors.
Genes 2019, 10(3), 222; https://doi.org/10.3390/genes10030222
Submission received: 31 January 2019 / Revised: 1 March 2019 / Accepted: 12 March 2019 / Published: 15 March 2019
(This article belongs to the Special Issue Genetic Determinants of Human Longevity)

Abstract

:
Human longevity is a complex phenotype resulting from the combinations of context-dependent gene-environment interactions that require analysis as a dynamic process in a cohesive ecological and evolutionary framework. Genome-wide association (GWAS) and whole-genome sequencing (WGS) studies on centenarians pointed toward the inclusion of the apolipoprotein E (APOE) polymorphisms ε2 and ε4, as implicated in the attainment of extreme longevity, which refers to their effect in age-related Alzheimer’s disease (AD) and cardiovascular disease (CVD). In this case, the available literature on APOE and its involvement in longevity is described according to an anthropological and population genetics perspective. This aims to highlight the evolutionary history of this gene, how its participation in several biological pathways relates to human longevity, and which evolutionary dynamics may have shaped the distribution of APOE haplotypes across the globe. Its potential adaptive role will be described along with implications for the study of longevity in different human groups. This review also presents an updated overview of the worldwide distribution of APOE alleles based on modern day data from public databases and ancient DNA samples retrieved from literature in the attempt to understand the spatial and temporal frame in which present-day patterns of APOE variation evolved.

1. Introduction

The study of APOE and its isoforms has spread in all the studies about the genetics of human longevity and this is one of the first genes that emerged in candidate-gene studies and in genome-wide analysis in different human populations. The pleiotropic roles of this gene as well as the pattern of variability across different human groups provide an interesting perspective on the analysis of the evolutionary relationship between human genetics, environmental variables, and the attainment of extreme longevity as a healthy phenotype. In the present review, the following topics will be discussed.
  • APOE gene and protein structure and function, including the latest theoretical models describing its mechanism of action
  • The role of APOE in human longevity, its physiological functions, and the involvement in pathological traits in modern populations
  • APOE evolution and variability among human populations, including a novel analysis of modern and ancient data
  • The evolutionary mechanisms that maintained APOE deleterious variants in modern human populations.

2. APOE Structure and Models

Human APOE is a 299-amino acid long protein (34 kDa in weight) belonging to the family of amphiphilic exchangeable apolipo-proteins that is expressed in hepatocytes, monocytes/macrophages, adipocytes, astrocytes, and kidney cells [1,2,3,4]. Structural studies have shown two independently-folded domains for the lipid-free protein: an N-terminal elongated domain (residues 1–167) forms a 4 α-helix cluster in which non-polar residues face the inside of the protein, while the C-terminal domain (residues 206–299) has a more relaxed structure, with α-helices generating a largely exposed hydrophobic surface [5,6]. These domains are connected by an unstructured hinge that provides a large degree of mobility, which is necessary for the protein to fulfill its primary function in the hepatic and extra-hepatic uptake of plasma lipoprotein and cholesterol [7].
The N-terminal domain contains the low-density-lipoprotein receptor (LDLR) binding region, which is a cluster of basic arginine and lysine residues, spanning between positions 135 and 150 in helix 4 (an Arg-172 residue in the hinge is also necessary for the binding function [8]).A stretch of hydrophobic residues at the end of the C-terminal domain (residues 260–299) is deemed to be responsible for binding the protein to lipids as well as for directing oligomerization of lipid-free ApoE. Since the monomer is the form that binds to lipids, oligomer dissociation appears to be the rate-limiting step of protein lipidation [9,10].
The gene itself is located on chromosome 19:q13.3, together with the apoC genes APOC1, APOC2, and APOC4, which are members of the exchangeable lipoprotein family, and in proximity to the mitochondrial translocase of the outer membrane gene (TOMM40). This is another locus involved in the development of AD [11,12,13,14,15].
As represented in Figure 1, the combination of two mutations at the APOE gene (rs7412 C/T and rs429358 C/T) gives rise to the three main protein variants, called ε2, ε3, and ε4 (or, alternatively, APOE2, APOE3 and APOE4) [16,17,18]. Isoform ε3 has a cysteine in position 112 and an arginine residue in position 158, while isoform ε2 has two cysteine residues and isoform ε4 has two arginine residues. Several other mutations can act on this background to nuance the effects of the three main variants and are involved in diverse cardiovascular pathologies, as reported, for example, in a recent review by Matsunaga and Saito [19].
While the difference in sequence is limited to a couple of residues, this has a great impact on the protein biophysical and, consequently, functional properties, since the change in structural features of APOE provides insight on the different behavior of its isoforms [20,21,22,23,24,25,26].
In particular, the Arg158Cys mutation in isoform ε2 reduces the affinity of the protein for the LDLR 50-to-100-fold [27] due to the removal of a crucial electrostatic interaction with Asp154. Mutating this residue to a neutral alanine has shown that the isoform fully recovers its functionality [28].
The mutation Cys112Arg in isoform ε4 does not change its affinity for the receptor but its preference for lipoprotein binding shifts from HDL (as do ε3 and ε2) to LDL/VLDL. This occurs because charged residues that should be buried in the protein core are, instead, propelled outwards and can establish trans-domain interactions that modify the protein structure and, therefore, lipoprotein preference, possibly by hindering overall dynamics [29,30]. Mutagenesis experiments proved effective in re-establishing the preference of isoform ε4 for HDL [17,29,31,32].
Both domain interactions and intermolecular interactions have been recently confirmed by using Forster Resonance Energy Transfer assay (FRET), which is a method to quantify the exchange of energy between two fluorescent tags attached to the ends of the APOE protein. These experiments showed that there is a consistently significant difference among isoforms, with ε4 showing a higher degree of energy transfer for both domain interaction and polymerization. However, a different study asserted that conformational changes appeared to reduce the propensity of this isoform to self-stabilize in tetramers [33,34].
Denaturation experiments aimed at testing protein stability again showed different behaviors for the three isoforms, with the ε2 N-terminal domain being the most resistant and being followed by ε3 and ε4, which is the least resistant isoform, but shows a higher number of stable intermediates between its folded and unfolded forms [35,36,37,38,39]. This has been interpreted as isoform ε4 assuming partially unfolded stable states at different pH in basic environments, facilitating large conformational changes and, in doing so, increasing the remodeling rate of lipoprotein particles. This has also been noted with other exchangeable apolipo-proteins, such as APOAI and APOAII [38,39,40,41]. Higher ε4 catabolism, although being not an index of overall increased efficiency in plasma lipoprotein clearance, may justify why APOE4 homozygotes have a lower plasma APOE concentration [42,43,44,45]. On the other hand, it has been suggested that partially folded APOE is more sensitive to proteolysis of the domain-connecting hinge and that isoform ε4 may be more easily flagged as “misfolded” due to domain interaction, particularly in the brain [46,47,48,49,50].
It is also important to remember that no definitive mechanism for how APOE binds to lipids has been elucidated even though different hypotheses have emerged over the years, especially in relation to the implication of its isoforms in pathological traits. Starting from the concept of “molten globule” [36,51], a hairpin model has been proposed assuming that the protein bends itself so that the LDLR-binding motif is exposed at one extremity of the structure [31,52,53,54,55]. Other studies have suggested a conformational heterogeneity of bound apoE, observing that LDLR binding affinity, while higher in the bound protein than in the lipid-free protein, is modulated by the particle size, its lipid composition, and the presence of other bound lipoproteins [31,52,56,57,58].
A revised model has been recently proposed and considers the high proportion of intrinsically disordered regions in the protein (up to a third of the whole molecular structure), multiple interactions between the two domains, the presence of evolutionarily conserved residues, and structural differences that may justify the lipid-binding preferences of isoforms ε3 and ε4 [20,59]. The authors of this work also argue that most structural studies on lipid-bound apoE make use of the hepatocyte-secreted protein and plasma lipids, but that the lipid composition in the brain is different and the current models may fail to address lipidation mechanisms of astrocyte-synthesized APOE [59].

3. APOE Function and Pathology

Multiple lipid-related physiological functions are associated with APOE. In particular, isoform ε3 helps in maintaining the structural integrity of cholesterol-rich lipoproteins and enhances their solubilization in blood plasma, regulates lipid homeostasis of both hepatic and non-hepatic tissues, facilitates lipid internalization in cells and, when expressed by lipid-laden macrophages after cellular clearance, activates the reverse cholesterol transport, redirecting any excess of cholesterol to the liver for elimination [60,61,62,63].
The APOE genotype accounts for the vast majority of AD risk and AD pathology: inheriting one copy of APOE4 raises a person’s risk of developing the disease fourfold, while, with two copies, the risk increases 12-fold [64]. Raber and colleagues and, at the same time, Saunders and colleagues reported that clinical data regarding the association of the ε4 allele with AD suggests that 50% of AD is associated with the ε4 allele in the United States [65,66]. APOE4 may be responsible for the accelerated formation of β-pleated amyloid, as supported by studies showing that individuals with two copies of the APOE ε4 allele have a higher risk and earlier onset than heterozygous subject [67]. Moreover, a significant increase in risk of EOAD (early-onset Alzheimer’s disease) was found for individuals homozygous for APOE4 regardless of family history of dementia, but an increase in EOAD risk for APOE4 heterozygotes could only be shown in subjects with a positive family history [68].
Experiments with knock-out mice have proven that failed expression of APOE leads to a shortened lifespan due to the emergence of typically age-related phenotypes like an altered lipoprotein profile (the forefront of atherosclerosis and cardiovascular disease), neurological disorders, type II diabetes, deficits in immune response, and elevated markers of oxidative stress [69,70,71,72,73,74,75]. Moreover, the APOE variants determining the three isoforms ε2, ε3, and ε4 have also been associated with the modulation of body mass index (BMI) at statistical significance (p < 10−3) in a meta-analysis including 27,863 individuals from seven longitudinal cohort studies [76]. This highlights, on one hand, that APOE is a pleiotropic gene that simultaneously affects multiple phenotypes, depending on the site of protein synthesis (in particular, liver and brain). On the other hand, this emphasizes that the manifestations of its impairment fit the definition of aging as a general decline in biological functions, decreased stress resistance, and elevated susceptibility to disease that leads to an increase in mortality with age [77,78,79].
Most of the research conducted at this point focused on isoform ε4 as the “functionally altered” form of APOE in the brain since this is one of the most consistent candidates associated with human longevity and the onset of AD, according to GWAS and whole genome sequencing studies [62,66,68,80].
The finding of unexpectedly large proportions of C-terminal APOE in β-amyloid plaques of ε4/ε4 homozygous AD subjects leads to the hypothesis that the partially folded protein is highly sensitive to proteolysis [46,47,48,49,50] and this prevents APOE in helping Aβ clearance, favoring instead its deposition [81]. By folding into a more helical structure, truncated ε4-165 was shown to have deleterious effects on this same process, which stresses that structural integrity is important for AD pathogenesis [82,83,84]. The link with Aβ has also been associated with a higher degree of lysosome leakage in neurons, primarily due to the enhanced lipid remodeling activity of isoform ε4 on the lysosomal membrane at a low pH [85,86].
Experiments on mice have highlighted how isoform ε4 can also cause behavioral deficits in the absence of amyloid accumulation and, as with AD in humans, spatial and memory impairments increase with age and are observed primarily in females [87,88,89,90]. Regarding neuronal plasticity, similar studies showed that isoform ε3 associated with VLDL clearly stimulates neurite extension in developing neurons by feeding their membrane with lipids, while isoform ε4 inhibits branching likely due to effects on microtubule stability mediated by the LDLR-protein signaling pathway. The ε4 isoform also inhibits GABAergic input in newly formed neurons [91,92,93,94].
Furthermore, this isoform has been associated with decreased cerebral glucose metabolism that occurs even decades before the cognitive impairment becomes apparent, which suggests an interaction with the mitochondrial membrane and components of the respiratory complexes III and IV at very early stages of the disease [95,96,97,98,99,100]. An interesting observation is that mitochondria and the endoplasmic reticulum (ER) are intimately connected via mitochondria-associated membranes (MAMs) and the protein miofusin-2, so that mitochondrial dysfunction may propagate to the ER and affect the secretory pathway [12,101]. If the protein is recognized as unfolded, the pathways of the unfolded protein response can activate an inflammatory process by stimulating NF-kB, which is a transcription and cytokine regulator that mediates the immune response in cell survival [102,103,104].
Isoform ε4 also shows a decrease in the anti-oxidative properties of APOE as a metal cation binding protein. In fact, APOE4 genotype correlates with a higher degree of lipid oxidation and presence of hydroxyl radical levels in the blood of post-mortem patients [71,105]. Macrophages overexpressing ε4 also display membrane oxidation and generate anion radicals and, as a stress response, an increase of the anti-inflammatory protein heme oxygenase 1, was observed [106].
Moreover, it has been noted that, because of the cholesterol binding property of APOE and the fact that cholesterol is the main component of the envelope of many human-infecting viruses, the different behaviors of isoforms ε3 and ε4 may, respectively, impede or ease infections. For example, extensive work in the last 20 years showed that herpes simplex virus HSV-1 is frequently found in the brain of elderly normal patients as well as AD-affected patients, and it is thought that isoform ε4 can facilitate the process of colonization and repeated activation of latent colonies through inflammation, which exacerbates neural decay at a younger age. It is also suggested that an antiviral therapy may be effective in slowing AD progression (see comprehensive reviews in References [107,108,109]). The hepatitis C virus, on the other hand, needs APOE for assembling and the host lipid metabolism is directly involved in the viral infection [110,111,112,113,114,115]. Lastly, an interesting set of studies tried to investigate a link between APOE and the modulation of HIV infection as a chronic disease, now that the affected individuals can live to older ages thanks to anti-retroviral therapy. Even though the overall results are somewhat contrasting, isoform ε4 seems to correlate in different cases with the development of HIV-associated neurocognitive disorders, impaired cognition, dyslipidaemia, premature brain aging, and increased chance of debilitating opportunistic infections [116,117,118,119,120] (see also a comprehensive review in Reference [121]).
However, one of the most notable associations to be examined is between APOE alleles and cardiovascular disease (CVD). A study carried out on nine cohorts (eight of European and one of Chinese ancestry) of middle-aged men recruited by the World Health Organization MONICA (Monitoring of Trends and Determinants in Cardiovascular Disease) Project showed how variation in the relative frequency of the ɛ4 allele could predict 40% to 75% of the variation in coronary heart disease (CHD) fatalities among populations and how a 0.01 increase in the frequency of this allele could increase CHD death rates by 24.5/100,000 [122]. A study on follow-up data for almost 1000 Danish and Finnish heart attack survivors similarly denoted that carrying this variant can be a prognostic element, as these subjects have an 80% increased risk of dying [123]. A similar conclusion is presented by a post-mortem study, performed at the Oslo University Hospital, on over 1500 individuals who died of natural causes. In the cohort of patients presenting a cardiovascular disease (35% of the total), there were significantly more ɛ4 carriers (34% against 29%) and significantly less ɛ2 carriers (12% against 14%) than in the rest of the group (p < 0.05) [124]. It has also been recently recognized that, not only APOE is associated to cardiovascular risk, but also with the level of unsaturated and saturated circulating fatty acids, so that some light is being shed on how environmental and dietary factors can mediate the association between APOE variants and adverse cardiovascular events [125].
The common APOE alleles ɛ2, ɛ3, and ɛ4 are located in a CpG island and the related SNPs impact on the quantity of CpG dinucleotide, which impacts the gene DNA methylation. A recent study showed that the DNA methylation profile of this genomic region differentiates AD brain if compared to that of control subjects [126]. Moreover, a recent study on lymphocytes showed that DNA methylation in the APOE gene is associated with age and shaped by genetic variants in the gene [114]. A different study in African Americans also suggested that DNA methylation in blood cells may be an early indicator of individuals at risk for dementia [127].

4. APOE and Human Longevity

Many studies have attempted to grasp the complexity of the genetics of human longevity [128,129,130,131,132,133]: recent findings suggest that alleles associated with this phenotype are population-specific and, at the same time, that the achievement of extreme longevity is modulated by mechanisms shared among populations [134,135,136]. One of the most relevant loci identified by many studies (if not all) is the APOE gene.
Candidate gene studies, genome-wide association studies (GWAS) on geographically diverse populations, and, more recently, whole-genome sequencing approaches have been aimed at uncovering the genetic variants that influence the longevity phenotype and APOE possibly due to its involvement in several post-reproductive pathologies, which has emerged as a strong candidate in most of them. In this section, a brief overview of the studies on human longevity conducted in relation to the three main variants of APOE is presented, with special attention to its isoforms ε2, ε3, and ε4 arising from the combination of two mutations (rs7412 C/T and rs429358 C/T) [16,17,18].
Several GWAS supported the association between APOE and the longevity trait. For example, a Japanese study including 743 centenarians and 822 middle-aged controls found a novel positive association between variant rs16835198-G of the gene FNDC5 (which synthetizes a pro-hormone that is upregulated by muscular exercise) and APOE alleles in individuals with extreme longevity, which further highlights the polygenic nature of this trait [137]. A recent meta-analysis of GWAS examined data from 6036 individuals at least 90 years old against a control group of 3757 subjects that died between the ages of 55 and 80. A replication of known variants at APOE and FOXO3 genes was obtained, but the authors also pointed out the difficulty in locating new alleles associated with survival past the age of 90, possibly because of heterogeneous genetic influences combined with the fact that rare variants are not usually picked up by GWAS [80]. A novel statistical method for evaluating genome-wide associations starting from previous knowledge of age-dependent and disease-related traits that overlap with longevity (i.e., informed GWAS, or iGWAS) was applied to reduce the background SNPs possibly associated with extreme ages and to amplify potential signals that could be difficult to pick up in small centenarian cohorts [138]. Accordingly, 92 SNPs at eight independent loci (including the APOE/TOMM40 locus) were found to be associated with longevity at GWAS significance (p < 10−8) and four of these were further replicated in three different validation cohorts including the APOE/TOMM40 rs4420638 variant [138].
However, other studies failed to identify significant associations. For example, a study involving a Chinese cohort of 312 individuals with at least one long-lived parent (i.e., aged over 90) and 298 controls without a familial history of longevity found no significant correlation between APOE isoforms, age, and the levels of blood cholesterol (HDL-C) even though HDL-C levels themselves are significantly higher in the longevity group (p = 0.0001) [139]. The first study on a Brazilian cohort, including 220 individuals of at least 85 years of age and 232 controls averaging 72 years, was recently performed to investigate the association between FOXO3, SOD2, SIRT1, and APOE known variants and several phenotypes in oldest-olds. Only an association of two FOXO3 alleles with gender and triglyceride levels was confirmed in this case and the authors suggested expansion of the number of samples in order to perform a more powerful analysis [140].
A similar pattern emerged from candidate gene studies, as some have highlighted putative associations between APOE and extreme lifespan, while others have not. For example, a study focused on three independent cohorts of centenarians from Italy, Spain, and Japan compared with healthy, younger controls confirmed the ε4 allele being negatively associated with extreme longevity in all three cases after adjustment for sex, while allele ε2 was positively associated with the same trait in the Japanese and Italian cohorts only. This highlighted that the ε4 variant appears to decrease the likelihood of reaching extreme ages across ethnicity and geographic origin [141]. A recently published paper on 450 individuals of Ashkenazi Jewish ancestry at least 95 years of age contrasted with 500 controls without a history of familial longevity, which undertook a full analysis of the coding and regulatory regions of APOE. Two common regulatory variants were, thus, found in the proximal promoter of the gene (rs405509 and rs769449), which is significantly depleted in the elderly group (p < 0.036). Moreover, a significant enrichment of the ε2 allele (p = 0.003) and the ε2/ε3 genotype (p = 0.005), as well as a reduction of the ε3/ε4 genotype (p = 0.005) were observed in the same group [142]. Two recent reviews and meta-analyses of polymorphisms associated with human longevity recovered genomic data of European and Asiatic cohorts involving centenarians (i.e., 13 cohorts [141,143,144,145,146,147,148,149,150,151,152,153] for the 2014 review [154], 12 cohort s [141,143,144,148,149,155,156,157,158] for the 2018 review [130]), and added newly generated data to obtain groups of at least 2700 centenarian cases and 11,000 younger controls. The first study highlighted how the likelihood of reaching extreme longevity is negatively associated with carrying the ε4 allele, the ε4/ε4, ε3/ε4, or ε2/ε4 genotypes (all p < 0.001), while the trait is positively associated with the ε2/ε3 genotype (p = 0.017) [154]. The second study ascertained the homogeneity between the European and Asian groups when accounting for ethnicity. It also confirmed a significant negative association of the ε4 allele with longevity and a positive association of the ε2 variant with the same trait (which was not supported by the 2014 meta-analysis [154]) when compared to the ε3 allele (p < 0.0001) [130]. In another meta-analysis, data of over 28,000 individuals born between 1880 and 1975 were collected from seven studies on population longevity and familial healthy aging, with cases ranging from 96 to 119 years and controls from 0 to 99 years. Three genetic models (i.e., standard genotypic model, additive model for the effects of the ε2 allele, grouping of genotypes containing and not containing ε4) and two definitions of longevity (i.e., age at death, age reached by less than 1% of the population) were applied. Results showed that carrying the ε2 allele, but not ε4, is associated with significantly increased odds of reaching extreme longevity, with decreased risk of death when compared to the most common genotype ε3/ε3, but modest risk reduction at the most extreme ages. The opposite is observed for ε4, which acts independently from ε2 and associates with decreased odds for extended lifespan and an increased death risk that persists even at extreme ages in all groups. Furthermore, a joint haplotype analysis of five SNPs at the PRVL2-TOMM40-APOE-APOC1 gene cluster revealed that three haplotypes were individually associated with extreme lifespan when compared to the most common haplotype. The first one, containing ε2, was associated with a 34% increase in odds of extreme longevity (p = 7.8 × 10−7). The second one, containing ε4, was associated with a 50% decrease in the same odds (p = 10−8). The last one was, instead, an uncommon haplotype containing ε3 and was associated with a 20% decrease in odds for extreme longevity (p = 0.04), which suggests that there are SNPs at this locus that can exert a negative effect on longevity independently from the influence of the APOE ε4 allele [159].
A more extensive collection of GWAS and candidate gene studies performed in the last 8 years and describing APOE gene variants in human longevity is reported in Supplementary Table S1.
A recently published paper about genetic variants affecting viability over generations in large cohorts applied a method for testing the variability in allele frequency across different ages, after considering individual ancestry. When applied to the Genetic Epidemiology Research in Adult health and Aging (GERA) cohort and to parents of the UK Biobank participants, few common variants significantly related to mortality at specific ages were found across the genome, all tagging the APOE ε4 allele and the CHRNA3 gene. When testing for viability effects of genetic variant sets, strong signals (p < 10−3) were found relating delayed puberty with longer parental lifespan, as well as later age of first birth with longer maternal lifespan and, lastly, cholesterol levels and risk of coronary artery disease, with a marked difference between male and female participants [160].
It is worth noting that recent data from Northern European populations [148,161] clarified that APOE variation is associated with the likelihood of reaching extreme longevity not because it is a ‘longevity gene’ that ‘ensures’ a long life by itself, but due to the fact that it is rather a ‘frailty gene’ that slightly influences mortality and, particularly, ε4 is associated with an increased risk for death that persists even beyond ages reached by less than 1% of the population [159].

5. APOE Evolution and Variability among Human Populations

Human APOE clusters with members of the groups APOA and APOC in the superfamily of exchangeable apolipoproteins. These are structurally and functionally distinct from the non-exchangeable apolipoproteins APOB48 and APOB100, which make up the core of the lipoprotein particles [162,163].
Phylogenetic reconstruction using apolipoprotein sequences from representative eukaryotic species has shown that an ancestral form of this protein already existed before Metazoan evolution (i.e., approximately 750 Mya) and that divergence between the exchangeable and non-exchangeable families is equally ancient [162]. Focusing on the human exchangeable superfamily, a similar analysis showed that APOE clusters specifically with APOA1, APOA4, and APOA5 (the most recently identified human apolipoprotein), are separated from the cluster including APOA2, APOC3, APOC2, and APOC1 (the oldest in the cluster). It is also noteworthy that the length of the synthetized protein increases from the oldest to the youngest gene [162]. When including the insect apolipo-protein ApoLpIII in the analysis, it was found to group by sequence similarity within the human APOE cluster, instead of being an outgroup to all human exchangeable proteins. This suggests that the divergence of exchangeable apolipo-proteins occurred at an early evolutionary stage, possibly with the advent of bilateral symmetry (i.e., approximately 650 Mya), while the origin of ApoLpIII has dated back to the emergence of flying insects (i.e., 500 Mya) [162,164,165]. Nevertheless, an extensive review of phylogenetic relationships among eukaryotic apolipo-proteins is not the purpose of this review [162].
Focusing on the investigation of human-specific apolipo-proteins characteristics, comparison of the protein sequence of human and primate APOE reveals that the non-human apolipo-protein has arginine in position 112, like human isoform ε4. This suggests that ε4 is the ancestral variant and recent analyses of Denisovan DNA (a specimen of archaic human found in 2010 in the Denisova cave, in the Altai Mountains in Siberia) also corroborate such a hypothesis [166,167]. Unfortunately, this information is not yet fully disentangled for the Neanderthal genomes. The other non-synonymous variants detected among the species do not alter the size or charge of the residues and are not located in functional domains [162]. The only fundamental difference, then, involves residue 61, where humans present an arginine, while all other primates have a threonine. The Thr61Arg substitution introduces a bulkier, positively charged residue near the equally charged Arg112, by which it is projected out of the N-terminal helix bundle. This repositioning allows for Arg61 in ε4 to be involved in domain interactions that affect the isoform structure, which makes the protein less stable, but readier in binding large, lipid-rich lipoproteins. It is, however, unclear how the mutation that originated human ε4 from an ancestral APOE could provide a net evolutionary advantage. Theories including the consumption of cholesterol-rich meat, the presence of pathogens in uncooked foods, and increasing brain size during human evolution have been proposed as well as random DNA photooxidation following the loss of body hair [162,168].
One of the most intriguing hypotheses for the development of longevity despite the presence of a deleterious APOE isoform, however, postulates a link with increased physical activity, over the evolutionary history of the genus Homo, that helped in counterbalancing a higher risk of cardiovascular disease [169]. Haplotype analysis revealed that the origin of isoforms ε2 and ε3 in humans can be dated back to 200,000 to 300,000 years ago [170], while the increase in physical exercise occurred much earlier in time, possibly around 1.8 Mya, when Homo erectus abandoned the sedentary lifestyle of the forests to become a hunter-gatherer. Long foraging distances and the ability to run for extended periods of time, to either follow prey or flee from danger, require endurance and increased levels of aerobic activity, which is related to the conversion of body fat into usable energy and is in stark contrast with the cardiovascular effects induced by the ε4/ε4 haplotype [171,172,173]. This likely relaxed the limitation on lifespan imposed by the deleterious allele and is in accordance with fossil dating and palaeodemographic analyses that testify an increase in the number of older individuals throughout the evolution of H. erectus and then H. sapiens [174], as well as the extension of post-reproductive lifespan in concert with the development of a hunter-gatherer lifestyle [169,175,176].
However, in modern populations, isoform ε4 is only the second-most common APOE variant, which shows the highest frequency in indigenous populations of Central Africa (40% in Aka Pygmies, 38% in Tutsis, 33% in Zairians, and 29% in Fon), Oceania (49% in the Hui population of New Guinea, 26% in the Mowanjum aboriginal tribe of Western Australia and in Polynesians from American Samoa) and Mexico (27% for the Huychol in Nayarit [177]). Isoform ε3, instead, shows peaks of 94% in the Alberta Hutterite people of Canada, 90% in Mexican Mayas, 88% in the Basque and Sardinian populations of Europe, and 86% in Han Chinese. As highlighted in Figure 2 and reported in Supplementary Table S2, a distinct latitudinal gradient for ε4 can be observed across Europe (5% to 10% in Spain, Portugal, Italy, and Greece, up to 16% in France, Belgium, and Germany, up to 23% in the Scandinavian peninsula, with peaks of 31% in the Saami population of Finland) and it has been also reported in China (5% to 17.5% in 19 distinct populations) [178,179,180,181]. In the context of the present review, data have been also gathered for a cohort of 134 Italian centenarians and 350 healthy, younger controls, so that 484 samples were enrolled in three Italian areas (North, Center, and South Italy) and clustered according to their place of birth. DNA samples were recovered after approval by the Ethical Committee of Sant’Orsola-Malpighi University Hospital (Bologna, Italy). As shown in Supplementary Tables S3 and S4, when individuals from both groups were separately clustered by macroareas [182,183], a definite gradient could be observed for the ε4 allele in both centenarians and controls, with frequencies of 0.125 and 0.124, respectively, in Northern Italy, 0.052 and 0.063 in Central Italy, and 0.026 and 0.039 in Southern Italy. Although sample size is relatively small in the latter group, the increase in frequency from South to North at both a regional and a continental level follows a pattern that has been already observed. For example, in Italy and in Europe, for other genes involved in lipid metabolism [182,184,185], this suggests that isoform ε3 may be selected against ε4 at lower latitudes, but this does not explain the evolutionary advantage of the single amino acidic mutation Arg112Cys provided in giving rise to the now most frequent APOE variant worldwide [178,179,180,186]. Studies on this topic report a higher structural stability and functional flexibility of isoform ε3, which can also be associated with metal binding, oxidative stress resistance, micronutrient uptake, enhanced neuronal repair following damage, and an energy-conserving phenotype [187,188,189,190,191] (see a comprehensive review on adaptation to dietary changes in Reference [192]). However, being more adaptive and responsive to environmental changes does not justify that all the ailments of isoform ε4 is associated with, tend to be post-reproductive. Theories have been recently introduced that several derived alleles (including those at the APOE gene) with a protective effect on cognition after menopause may result from late-life selection through an increase in younger kin survival. The proposal of this “grandmother effect” may explain the predominance of the ε3 allele in a trans-generational way by assessing that the extension of the post-reproductive lifespan as a healthy phenotype requires the prevention of age-related cognitive decline to increase the survival of younger kin under grandparental care. Moreover, cultural transmission through generations is known to shape the social structure of modern foraging populations, which enhances the survival probability of the individuals belonging to networks that are enriched in multi-generational sharing of knowledge [175,176,193,194].

6. APOE Trade-Offs

Human longevity is a complex phenotype in which small contributions from a high number of genetic variants participate to define most age-related traits in later life. Isoform ε4 of APOE is involved in several cardiovascular and neural pathologies that become apparent at a post-reproductive age. Many studies in the last decade tried to find explanations as to why such a deleterious variant has been maintained at high frequency in many human groups, particularly in indigenous populations of Africa and Oceania [178]. The main collected findings suggest an association between isoform ε4 and a number of population-specific and environment-related beneficial effects that compensate for the damage induced by the same variant in later life [175,176,187,189,193].
The observation that the most detrimental effects of APOE (CVD, AD, reduced lifespan) mainly affect individuals of affluent populations, while most African groups do not develop significant impairment despite presenting the highest frequencies of isoform ε4, prompted a study on a rural Ghanaian population characterized by high levels of mortality from widespread infectious diseases. The analyses conducted pinpointed an association between the exposure of fertile women to high pathogen levels and a higher degree of fertility (ε4 carriers have one more child than non-carriers, while ε4 homozygous women have 3.5 more children on average). Such polymorphism may be maintained because it favors reproduction in a context where limited survival at older ages spontaneously delays the detrimental effects of the isoform. Conversely, individuals living in modernized societies, less affected by pathogens and capable of reaching an older age, have no need for the positive reproductive advantage conferred by this allele and have, instead, more probability to manifest the related negative repercussions [195]. Moreover, several candidate gene studies conducted on cohorts from industrialized countries (e.g., Iran, Turkey, United States) seem to highlight a positive relationship between cardiovascular disease and thrombophilia, as induced by lipid clearance dysfunction through the ε4 variant, and occurrences of two or more consecutive miscarriages before the 20th week of gestation. These studies compared groups of affected women and fertile negative controls with at least two successful pregnancies. In all cases, a statistically significant enrichment in the ε4 variant was found for the cohorts affected by recurrent pregnancy loss as well as a significant positive association of the ε3/ε4 and ε4/ε4 genotypes with the analyzed phenotype [196,197,198,199,200].
Another study relating pathogen exposure to the preservation of the deleterious isoform was performed on the Tsimane population of Amazonian foragers. Results highlighted that ε4 carriers with a high eosinophil count (a sign of parasitic infection) perform better in cognitive tests than the non-infected carriers, irrespective of their age [201,202].
Some publications also support the thesis that the extremely long span of human survival beyond fertile age is an exception in the world of primates and mammals and is tightly linked to the practice of inter-generational cooperative child rearing, which potentially developed early in hunter-gatherer societies. The role of the grandmother is, in this case, equally parted in practices of active support, information transfer, and building of social networks that can result in extensive sharing of resources, which favor the survival and growth of the younger individuals. In this case, the positive effects of differential survival and reproductive success in early life are mirrored by deleterious cognitive deficiencies at an older age, when natural selection is absent [175,193].
Other studies have proposed that the main advantage provided by isoform ε3 when it first emerged, around 200,000 years ago, relates to an early shift in dietary habits. More organized hunting methods and the use of fire enhanced the quantity of fat-rich meat introduced with diet, which ultimately helped extend the human lifespan. Survival to reproductive age and beyond would, in this case, require both an efficient clearance of excess cholesterol from the blood and a stronger inflammatory response to food-borne pathogens, which is provided by the more ancient isoform ε4 [168,192].
The ε4 allele is an independent risk factor in age-related mortality and all-cause mortality. Since it hampers longevity, one would expect a general reduction of allele frequency with increasing age. However, the disease risk association seems to vary in an ethnic-related way. For example, hypertension and brain hemorrhage risks are increased only in Asian and European ε4 carriers [203,204], while African and Hispanic Americans show an increased risk for Alzheimer disease even in the absence of ε4, which allows for its accumulation in older age cohorts, because it is less detrimental [178,205]. Other studies have shown how this variant may exert negative pleiotropy, which grants protection to the infant brain and against infections at a younger age. This counterbalances the deleterious effects that may be induced later in life [206,207,208,209].
Lastly, isoform ε2 has a worldwide frequency of around 7% and a patchy distribution, with peaks in Southeast Asia, Australia, and some African populations (up to 19%) and absence in most indigenous American groups [178]. The effects of this isoform are opposite to those of ε4. Carriers show a lower risk and delayed onset of cognitive decline and a significantly reduced risk of cardiovascular disease, but increased infection rates at a young age [208,210,211,212,213,214]. Given the opposite effects of the two isoforms, the only current explanation for their simultaneous high frequency in several indigenous African populations is that selection acts for ε2 and ε4 against ε3, but no definitive selective mechanism has been described so far [162].
Other possible explanations for the latitudinal distribution of APOE variants and the maintenance of ε4 relate to its role in immunity. As highlighted by ecological and biogeographical research, there is a clear relationship between the current distribution of human infectious diseases, latitude, average temperature, humidity, and population density, with harmful bacteria flourishing in hot, wet climates and in densely populated areas of the world [215,216]. Studies involving knock-out chimeras in mice suggest that APOE deficiency (also mimicking reduced functionality of ε4) leads to cholesterol build-up in dendritic cell membranes, which enhances antigen presentation via lipid rafts and increasing T-cell activity (hampering macrophage function [106,217]) regardless of ensuing hypercholesterolemia. This has also been directly observed in humans, where subjects expressing the ε4 isoform have a higher activated T-cell count when compared to carriers of the other isoforms [218]. However, earlier studies on mice also highlighted that APOE-deficient specimens may show a significantly reduced immune response to specific pathogens by becoming more susceptible to Lysteria monocytogenes and Klebsiella pneumoniae infections [72,73,219]. As described in paragraph 3.2, several viruses require the most common form of APOE to build their particles and invade human cells. In fact, it has been observed that isoform ε4 may hamper virion synthesis and compete with the hepatitis C virus for access to LDL receptors, which reduces liver damage in exposed populations. For example, in the Italian peninsula, the North-South gradient of hepatitis C incidence overlaps with a reverse gradient in ε4 distribution [220,221]. It is less clear how the different isoforms of APOE interact with the herpes simplex virus and HIV even though the mechanisms proposed in a review by Kuhlman et al. (2010) suggest that ε4, in this case, poses less competition to cell entry, which is also helped by the enhanced presence of lipid rafts in the cell membrane [222]. A relationship between APOE4 conservation, enhanced immune response, and pathogen distribution can be further justified by studies highlighting how carriers of this allele show higher levels of the anti-parasitic cytokine interleukin-3 (IL-3) and the pro-inflammatory tumor necrosis factor (TNFα) when exposed to endotoxins [223,224]. This seems to be especially important in the extreme case of Gram-negative infections since their toxins are membrane lipopolysaccharides (LPS) that can be collected by lipoproteins and redirected by APOE to the liver for inactivation. Reduced functionality of this protein can, thus, lead to hampered endotoxin clearance, overstimulation of macrophages, overproduction of inflammatory cytokines, and a stronger immune response leading to sepsis in the afflicted subject [225].
While local accumulation of the ε4 isoform in indigenous populations can be justified by the prevalence of infections in the absence of medical care, it can also be associated with a stronger inflammatory response to food-borne pathogens [168,192]. Other dietary factors, such as vitamin D and bone calcium assimilation, which were proven to be higher in both humans and transgenic mice carrying the ε4 allele [188,191], may have been crucial in the adaptation of populations living at extreme latitudes to the reduced amount of UV radiation. This justifies the North-South distribution of ε4 observable in Europe [188].
Many recent studies also considered a relationship between APOE and the gut microbiota, since, in this context, APOE can simultaneously exert its double role in lipid assimilation and immunity. Several experiments using APOE knock-out mice have shown that the diet can modulate gut microbiota composition such as with an enrichment in Firmicutes when mice were fed a typically Western diet. In turn, this relates to the amount of metabolic endotoxins in the bloodstream that stimulated a chronic inflammatory state [226,227]. On the other hand, if mice feeding on a hyperlipidic diet were immunized against their own gut microbiota, a significant decrement in serum inflammatory cytokines could be observed together with a reduction in atherosclerotic plaques, which suggests an interesting trade-off mechanism that balances the immune response against the resident microbiota with immune regulation of inflammation mediated by apolipoprotein E [228]. Other studies on obese mice and knock-out mice fed on regular chow versus a Western diet discovered that mending the loss of specific bacteria strains (e.g., Akkermansia muciniphila) caused by a hyperlipidic diet contrasted the enhanced permeability of the gastrointestinal tract to endotoxins and reduced vessel inflammation, fat dysmetabolism, and atherosclerosis both in normal and obese specimens [229,230,231]. Taking into consideration the immunomodulatory function of APOE, not only against bacteria, but also toward oxidized LDL found in sclerotic vessels, these observations highlight how both local and systemic responses can shape the overall arrangement of the intestinal biome [228].
Trade-off mechanisms may explain, in certain cases, issues regarding the replication of association signals for the same allele in different human populations and that several studies deem it more likely that a proportion of genetic influence on longevity (and of complex traits in general) may be explained through polygenic effects [232,233,234]. Furthermore, the studies performed until now did not fully address the role of rare mutations [235] nor the interaction between rare variants and APOE that seems to have a relevant impact on the phenotypic outcome, as supported by a recent study on the Hong Kong Chinese population [236]. Lastly, in this review, we did not address a potential limitation of trade-off mechanisms: the fact that they may be time-dependent and may be influenced by specific environmental (internal and external) conditions.
The contrast between APOE4 and APOE3 frequency distributions in current populations, with the former being prevalent in foraging communities and the latter being predominant in regions with relevant agricultural economy, led to the theory that the ε4 variant is a relic of a hunter-gatherer genetic background that has not adapted to the modern, energy-rich, and exercise-poor lifestyle [237]. To assess the possibility of observing the temporal scale of this transition, in the context of the present review, we built a panel of 1149 publicly available ancient genomes and selected 97 of them, with both rs7412 and rs429358 already directly genotyped (the original works including the selected samples can be found at References [238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258]). This has been done in order to avoid the introduction of bias in the dataset by imputing variants from highly deteriorated DNA, which usually presents extended regions of missing data. The samples, mapped in Figure 3 and listed in Supplementary Table S5 with details on the place of discovery and cultural context, cover the Euro-Mediterranean area and range from 1500 to 42,000 years ago. The ε3/ε3 genotype was found to be the most frequent (83%), followed by the ε4/ε4 genotype (13%), and the ε2/ε2 genotype (3%). The only heterozygote ε3/ε4 was represented by the Ust’Ishim sample, a 42,000-year old specimen of early hunter-gatherer human found in Siberia. In more detail, the ε2/ε2 individuals are Northern European samples from the Bronze Age. Despite carrying the ancestral genotype, all ε4/ε4 individuals are less than 8000 years old, with most of them being even more recent than 5000 years, while a conspicuous number of ε3/ε3 samples are much older than this, especially in the areas of Caucasus, between the Black Sea, the Caspian Sea, and the Middle East. This temporal and spatial distribution may be coherent with Palaeolithic alleles, like APOE4, having been reintroduced in Europe at higher frequency with the Yamnaya migration from the Steppe during the Bronze age and APOE3 being present at higher frequencies in the Fertile Crescent prior to the Neolithic Revolution, even though both alleles were already present in the European populations as well, as highlighted by the older local specimens [238,243,245]. However, the limited number of samples available across such an extended geographic area and the chance of genotyping errors due to the highly deteriorated ancient DNA hinder the possibility of a thorough factual discussion of the results. In order to draw more elaborate conclusions, it would be useful to recover more complete and evenly distributed ancient data, both in space and in time.

7. Conclusions

This review reports and summarizes relevant considerations regarding APOE and its pivotal role in the genetics of human longevity. Both candidate-gene studies and genome-wide analysis reveal its involvement in the attainment of an extreme lifespan by exerting a pleiotropic effect in a polygenic context. In this review, some new data (on the geographic distribution of APOE isoforms ε2, ε3 and ε4 in centenarians and in healthy individuals from the Italian population and on public available dataset on ancient genomes) have also been considered and evaluated in the light of the most recent findings on this gene, with particular attention to the variability across human populations. In fact, the study of the variability across different human groups is crucial to understand the differences that can be observed in the association between this gene and longevity and age-related diseases. The patterns can be justified by considering the multitude of biological pathways this gene belongs to and the different environmental conditions human populations must deal with especially with regard to pathogen exposure and dietary changes. An evolutionary perspective is also crucial to understand the conservation and current worldwide distribution of APOE isoforms ε2, ε3, and ε4. New data regarding DNA methylation variability in different tissues will also help more clearly define the role of this gene. Moreover, the relation between population specific cultural/ecological traits and APOE variability (as well as other genes) are needed to disentangle the devious way from genotype to phenotype. Given the high amount of data available on this gene, we think that an evolutionary approach, such as the one proposed by evolutionary medicine [259,260,261,262,263], will help interpret and clarify the link between even distant (or apparently not connected) results for this gene in different populations.

Supplementary Materials

The following is available at https://www.mdpi.com/2073-4425/10/3/222/s1. Table S1: Summary of the studies published in the last eight years investigating the association between APOE variants and human longevity. Table S2: APOE allele frequencies in different human populations. Data used for Figure 2. Table S3: number of APOE alleles and haplotypes in geographically divided groups of 484 Italian centenarians and controls. Table S4: frequency of APOE alleles and haplotypes in geographically divided groups of 484 Italian centenarians and controls. Table S5: Summary of APOE haplotypes in ancient genomes. Data used for Figure 3.

Author Contributions

C.G., P.A., D.L., C.F., and P.G. involved in the study design. P.A., M.S., P.G., A.B., D.M., C.F., D.L., and C.G. performed the literature review. PA performed data analysis. P.A., M.S., P.G., A.B., D.M., C.F., D.L., and C.G. performed a biological interpretation. P.A. and C.G. wrote the first draft and all authors were involved in reviewing and editing. C.F. and D.M. provided data on the Italian population.

Funding

This study was supported by The European Union’s Seventh Framework Program to CF (grant number 602757, HUMAN), the European Union’s H2020 Project to CF and PG (grant number 634821, PROPAG-AGING), and the JPco-fuND to CF (ADAGE) supported this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Blue, M.L.; Williams, D.L.; Zucker, S.; Khan, S.A.; Blum, C.B. Apolipoprotein E synthesis in human kidney, adrenal gland, and liver. Proc. Natl. Acad. Sci. USA 1983, 80, 283–287. [Google Scholar] [CrossRef] [PubMed]
  2. Kockx, M.; Traini, M.; Kritharides, L. Cell-specific production, secretion, and function of apolipoprotein E. J. Mol. Med. 2018, 96, 361–371. [Google Scholar] [CrossRef] [PubMed]
  3. Tedla, N.; Glaros, E.N.; Brunk, U.T.; Jessup, W.; Garner, B. Heterogeneous expression of apolipoprotein-E by human macrophages. Immunology 2004, 113, 338–347. [Google Scholar] [CrossRef] [PubMed]
  4. Boyles, J.K.; Pitas, R.E.; Wilson, E.; Mahley, R.W.; Taylor, J.M. Apolipoprotein E associated with astrocytic glia of the central nervous system and with nonmyelinating glia of the peripheral nervous system. J. Clin. Investig. 1985, 76, 1501–1513. [Google Scholar] [CrossRef] [PubMed]
  5. Wetterau, J.R.; Aggerbeck, L.P.; Rall, S.C.; Weisgraber, K.H. Human apolipoprotein E3 in aqueous solution. I. Evidence for two structural domains. J. Biol. Chem. 1988, 263, 6240–6248. [Google Scholar] [PubMed]
  6. Wilson, C.; Wardell, M.R.; Weisgraber, K.H.; Mahley, R.W.; Agard, D.A. Three-dimensional structure of the LDL receptor-binding domain of human apolipoprotein E. Science 1991, 252, 1817–1822. [Google Scholar] [CrossRef]
  7. Mahley, R.W.; Innerarity, T.L.; Rall, S.C.; Weisgraber, K.H. Plasma lipoproteins: Apolipoprotein structure and function. J. Lipid Res. 1984, 25, 1277–1294. [Google Scholar]
  8. Morrow, J.A.; Arnold, K.S.; Dong, J.; Balestra, M.E.; Innerarity, T.L.; Weisgraber, K.H. Effect of arginine 172 on the binding of apolipoprotein E to the low density lipoprotein receptor. J. Biol. Chem. 2000, 275, 2576–2580. [Google Scholar] [CrossRef]
  9. Huang, R.Y.-C.; Garai, K.; Frieden, C.; Gross, M.L. Hydrogen/deuterium exchange and electron-transfer dissociation mass spectrometry determine the interface and dynamics of apolipoprotein E oligomerization. Biochemistry 2011, 50, 9273–9282. [Google Scholar] [CrossRef]
  10. Chou, C.-Y.; Lin, Y.-L.; Huang, Y.-C.; Sheu, S.-Y.; Lin, T.-H.; Tsay, H.-J.; Chang, G.-G.; Shiao, M.-S. Structural variation in human apolipoprotein E3 and E4: Secondary structure, tertiary structure, and size distribution. Biophys. J. 2005, 88, 455–466. [Google Scholar] [CrossRef]
  11. Subramanian, S.; Gottschalk, W.K.; Kim, S.Y.; Roses, A.D.; Chiba-Falek, O. The effects of PPARγ on the regulation of the TOMM40-APOE-C1 genes cluster. Biochim. Biophys. Acta 2017, 1863, 810–816. [Google Scholar] [CrossRef]
  12. Roses, A.; Sundseth, S.; Saunders, A.; Gottschalk, W.; Burns, D.; Lutz, M. Understanding the genetics of APOE and TOMM40 and role of mitochondrial structure and function in clinical pharmacology of Alzheimer’s disease. Alzheimers Dement. 2016, 12, 687–694. [Google Scholar] [CrossRef]
  13. Cervantes, S.; Samaranch, L.; Vidal-Taboada, J.M.; Lamet, I.; Bullido, M.J.; Frank-García, A.; Coria, F.; Lleó, A.; Clarimón, J.; Lorenzo, E.; et al. Genetic variation in APOE cluster region and Alzheimer’s disease risk. Neurobiol. Aging 2011, 32, 2107.e7–2107.e17. [Google Scholar] [CrossRef]
  14. Papaioannou, I.; Simons, J.P.; Owen, J.S. Targeted in situ gene correction of dysfunctional APOE alleles to produce atheroprotective plasma ApoE3 protein. Cardiol. Res. Pract. 2012, 2012, 148796. [Google Scholar] [CrossRef]
  15. Kulminski, A.M.; Huang, J.; Wang, J.; He, L.; Loika, Y.; Culminskaya, I. Apolipoprotein E region molecular signatures of Alzheimer’s disease. Aging Cell 2018, 17, e12779. [Google Scholar] [CrossRef]
  16. Weisgraber, K.H.; Rall, S.C.; Mahley, R.W. Human E apoprotein heterogeneity. Cysteine-arginine interchanges in the amino acid sequence of the apo-E isoforms. J. Biol. Chem. 1981, 256, 9077–9083. [Google Scholar]
  17. Weisgraber, K.H. Apolipoprotein E distribution among human plasma lipoproteins: Role of the cysteine-arginine interchange at residue 112. J. Lipid Res. 1990, 31, 1503–1511. [Google Scholar]
  18. Chetty, P.S.; Mayne, L.; Lund-Katz, S.; Englander, S.W.; Phillips, M.C. Helical structure, stability, and dynamics in human apolipoprotein E3 and E4 by hydrogen exchange and mass spectrometry. Proc. Natl. Acad. Sci. USA 2017, 114, 968–973. [Google Scholar] [CrossRef]
  19. Matsunaga, A.; Saito, T. Apolipoprotein E mutations: A comparison between lipoprotein glomerulopathy and type III hyperlipoproteinemia. Clin. Exp. Nephrol. 2014, 18, 220–224. [Google Scholar] [CrossRef]
  20. Frieden, C.; Garai, K. Concerning the structure of apoE: Structure of apoE. Protein Sci. 2013, 22, 1820–1825. [Google Scholar] [CrossRef]
  21. Huang, Y.; Mahley, R.W. Apolipoprotein E: Structure and function in lipid metabolism, neurobiology, and Alzheimer’s diseases. Neurobiol. Dis. 2014, 72, 3–12. [Google Scholar] [CrossRef]
  22. Mahley, R.W.; Weisgraber, K.H.; Huang, Y. Apolipoprotein E: Structure determines function, from atherosclerosis to Alzheimer’s disease to AIDS. J. Lipid Res. 2009, 50, S183–S188. [Google Scholar] [CrossRef]
  23. Nguyen, D.; Dhanasekaran, P.; Nickel, M.; Mizuguchi, C.; Watanabe, M.; Saito, H.; Phillips, M.C.; Lund-Katz, S. Influence of domain stability on the properties of human apolipoprotein E3 and E4 and mouse Apolipoprotein E. Biochemistry 2014, 53, 4025–4033. [Google Scholar] [CrossRef]
  24. Henry, N.; Krammer, E.-M.; Stengel, F.; Adams, Q.; Van Liefferinge, F.; Hubin, E.; Chaves, R.; Efremov, R.; Aebersold, R.; Vandenbussche, G.; et al. Lipidated apolipoprotein E4 structure and its receptor binding mechanism determined by a combined cross-linking coupled to mass spectrometry and molecular dynamics approach. PLoS Comput. Biol. 2018, 14, e1006165. [Google Scholar] [CrossRef]
  25. Nguyen, D.; Dhanasekaran, P.; Nickel, M.; Nakatani, R.; Saito, H.; Phillips, M.C.; Lund-Katz, S. Molecular basis for the differences in lipid and lipoprotein binding properties of human apolipoproteins E3 and E4. Biochemistry 2010, 49, 10881–10889. [Google Scholar] [CrossRef]
  26. Nguyen, D.; Dhanasekaran, P.; Phillips, M.C.; Lund-Katz, S. Molecular mechanism of apolipoprotein E binding to lipoprotein particles. Biochemistry 2009, 48, 3025–3032. [Google Scholar] [CrossRef]
  27. Weisgraber, K.H.; Innerarity, T.L.; Mahley, R.W. Abnormal lipoprotein receptor-binding activity of the human E apoprotein due to cysteine-arginine interchange at a single site. J. Biol. Chem. 1982, 257, 2518–2521. [Google Scholar] [PubMed]
  28. Dong, L.-M.; Parkin, S.; Trakhanov, S.D.; Rupp, B.; Simmons, T.; Arnold, K.S.; Newhouse, Y.M.; Innerarity, T.L.; Weisgraber, K.H. Novel mechanism for defective receptor binding of apolipoprotein E2 in type III hyperlipoproteinemia. Nat. Struct. Mol. Biol. 1996, 3, 718–722. [Google Scholar] [CrossRef]
  29. Dong, L.-M.; Weisgraber, K.H. Human Apolipoprotein E4 Domain Interaction arginine 61 and glutamic acid 255 interact to direct the preference for very low density lipoproteins. J. Biol. Chem. 1996, 271, 19053–19057. [Google Scholar] [CrossRef] [PubMed]
  30. Dong, L.M.; Wilson, C.; Wardell, M.R.; Simmons, T.; Mahley, R.W.; Weisgraber, K.H.; Agard, D.A. Human apolipoprotein E. Role of arginine 61 in mediating the lipoprotein preferences of the E3 and E4 isoforms. J. Biol. Chem. 1994, 269, 22358–22365. [Google Scholar] [PubMed]
  31. Hatters, D.M.; Peters-Libeu, C.A.; Weisgraber, K.H. Engineering conformational destabilization into mouse apolipoprotein E A model for a unique property of human apolipoprotein E4. J. Biol. Chem. 2005, 280, 26477–26482. [Google Scholar] [CrossRef]
  32. Hatters, D.M.; Budamagunta, M.S.; Voss, J.C.; Weisgraber, K.H. Modulation of apolipoprotein E structure by domain interaction differences in lipid-bound and lipid-free forms. J. Biol. Chem. 2005, 280, 34288–34295. [Google Scholar] [CrossRef]
  33. Kara, E.; Marks, J.D.; Fan, Z.; Klickstein, J.A.; Roe, A.D.; Krogh, K.A.; Wegmann, S.; Maesako, M.; Luo, C.C.; Mylvaganam, R.; et al. Isoform- and cell type-specific structure of apolipoprotein E lipoparticles as revealed by a novel Forster resonance energy transfer assay. J. Biol. Chem. 2017, 292, 14720–14729. [Google Scholar] [CrossRef]
  34. Xu, Q.; Brecht, W.J.; Weisgraber, K.H.; Mahley, R.W.; Huang, Y. Apolipoprotein E4 domain interaction occurs in living neuronal cells as determined by fluorescence resonance energy transfer. J. Biol. Chem. 2004, 279, 25511–25516. [Google Scholar] [CrossRef]
  35. Morrow, J.A.; Segall, M.L.; Lund-Katz, S.; Phillips, M.C.; Knapp, M.; Rupp, B.; Weisgraber, K.H. Differences in stability among the human apolipoprotein E isoforms determined by the amino-terminal domain. Biochemistry 2000, 39, 11657–11666. [Google Scholar] [CrossRef]
  36. Morrow, J.A.; Hatters, D.M.; Lu, B.; Höchtl, P.; Oberg, K.A.; Rupp, B.; Weisgraber, K.H. Apolipoprotein E4 forms a Molten Globule A potential basis for its association with disease. J. Biol. Chem. 2002, 277, 50380–50385. [Google Scholar] [CrossRef]
  37. Acharya, P.; Segall, M.L.; Zaiou, M.; Morrow, J.; Weisgraber, K.H.; Phillips, M.C.; Lund-Katz, S.; Snow, J. Comparison of the stabilities and unfolding pathways of human apolipoprotein E isoforms by differential scanning calorimetry and circular dichroism. Biochim. Biophys. Acta BBA Mol. Cell Biol. Lipids 2002, 1584, 9–19. [Google Scholar] [CrossRef]
  38. Bychkova, V.E.; Ptitsyn, O.B. Folding intermediates are involved in genetic diseases? FEBS Lett. 1995, 359, 6–8. [Google Scholar] [CrossRef]
  39. Ptitsyn, O.B.; Bychkova, V.E.; Uversky, V.N. Kinetic and equilibrium folding intermediates. Philos. Trans. R. Soc. Lond. B 1995, 348, 35–41. [Google Scholar]
  40. Gursky, O.; Atkinson, D. Thermal unfolding of human high-density apolipoprotein A-1: Implications for a lipid-free molten globular state. Proc. Natl. Acad. Sci. USA 1996, 93, 2991–2995. [Google Scholar] [CrossRef]
  41. Gursky, O.; Atkinson, D. High- and low-temperature unfolding of human high-density apolipoprotein A-2. Protein Sci. Publ. Protein Soc. 1996, 5, 1874–1882. [Google Scholar] [CrossRef] [PubMed]
  42. Haddy, N.; Bacquer, D.D.; Chemaly, M.M.; Maurice, M.; Ehnholm, C.; Evans, A.; Sans, S.; do Martins, M.C.; Backer, G.D.; Siest, G.; et al. The importance of plasma apolipoprotein E concentration in addition to its common polymorphism on inter-individual variation in lipid levels: Results from Apo Europe. Eur. J. Hum. Genet. 2002, 10, 841–850. [Google Scholar] [CrossRef] [PubMed]
  43. Simon, R.; Girod, M.; Fonbonne, C.; Salvador, A.; Clément, Y.; Lantéri, P.; Amouyel, P.; Lambert, J.C.; Lemoine, J. Total ApoE and ApoE4 isoform assays in an Alzheimer’s disease case-control study by targeted mass spectrometry (n = 669): A pilot assay for methionine-containing proteotypic peptides. Mol. Cell. Proteom. 2012, 11, 1389–1403. [Google Scholar] [CrossRef]
  44. Martínez-Morillo, E.; Hansson, O.; Atagi, Y.; Bu, G.; Minthon, L.; Diamandis, E.P.; Nielsen, H.M. Total apolipoprotein E levels and specific isoform composition in cerebrospinal fluid and plasma from Alzheimer’s disease patients and controls. Acta Neuropathol. 2014, 127, 633–643. [Google Scholar] [CrossRef] [PubMed]
  45. Rezeli, M.; Zetterberg, H.; Blennow, K.; Brinkmalm, A.; Laurell, T.; Hansson, O.; Marko-Varga, G. Quantification of total apolipoprotein E and its specific isoforms in cerebrospinal fluid and blood in Alzheimer’s disease and other neurodegenerative diseases. EuPA Open Proteomics 2015, 8, 137–143. [Google Scholar] [CrossRef]
  46. Brecht, W.J.; Harris, F.M.; Chang, S.; Tesseur, I.; Yu, G.-Q.; Xu, Q.; Fish, J.D.; Wyss-Coray, T.; Buttini, M.; Mucke, L.; et al. Neuron-specific apolipoprotein E4 proteolysis is associated with increased Tau phosphorylation in brains of transgenic mice. J. Neurosci. 2004, 24, 2527–2534. [Google Scholar] [CrossRef]
  47. Riddell, D.R.; Zhou, H.; Atchison, K.; Warwick, H.K.; Atkinson, P.J.; Jefferson, J.; Xu, L.; Aschmies, S.; Kirksey, Y.; Hu, Y.; et al. Impact of apolipoprotein E (ApoE) polymorphism on brain ApoE levels. J. Neurosci. 2008, 28, 11445–11453. [Google Scholar] [CrossRef] [PubMed]
  48. Elliott, D.A.; Tsoi, K.; Holinkova, S.; Chan, S.L.; Kim, W.S.; Halliday, G.M.; Rye, K.-A.; Garner, B. Isoform-specific proteolysis of apolipoprotein-E in the brain. Neurobiol. Aging 2011, 32, 257–271. [Google Scholar] [CrossRef]
  49. Williams II, B.; Convertino, M.; Das, J.; Dokholyan, N.V. ApoE4-specific misfolded intermediate identified by molecular dynamics simulations. PLoS Comput. Biol. 2015, 11, e1004359. [Google Scholar] [CrossRef]
  50. Love, J.E.; Day, R.J.; Gause, J.W.; Brown, R.J.; Pu, X.; Theis, D.I.; Caraway, C.A.; Poon, W.W.; Rahman, A.A.; Morrison, B.E.; et al. Nuclear uptake of an amino-terminal fragment of apolipoprotein E4 promotes cell death and localizes within microglia of the Alzheimer’s disease brain. Int. J. Physiol. Pathophysiol. Pharmacol. 2017, 9, 40–57. [Google Scholar] [PubMed]
  51. Yeh, Y.-Q.; Liao, K.-F.; Shih, O.; Shiu, Y.-J.; Wu, W.-R.; Su, C.-J.; Lin, P.-C.; Jeng, U.-S. Probing the acid-induced packing structure changes of the molten globule domains of a protein near equilibrium unfolding. J. Phys. Chem. Lett. 2017, 8, 470–477. [Google Scholar] [CrossRef]
  52. Fisher, C.A.; Narayanaswami, V.; Ryan, R.O. The lipid-associated conformation of the low density lipoprotein receptor binding domain of human apolipoprotein E. J. Biol. Chem. 2000, 275, 33601–33606. [Google Scholar] [CrossRef] [PubMed]
  53. Narayanaswami, V.; Maiorano, J.N.; Dhanasekaran, P.; Ryan, R.O.; Phillips, M.C.; Lund-Katz, S.; Davidson, W.S. Helix orientation of the functional domains in apolipoprotein E in discoidal high density lipoprotein particles. J. Biol. Chem. 2004, 279, 14273–14279. [Google Scholar] [CrossRef]
  54. Newhouse, Y.; Peters-Libeu, C.; Weisgraber, K.H. Crystallization and preliminary X-ray diffraction analysis of apolipoprotein E-containing lipoprotein particles. Acta Crystallograph. Sect. F Struct. Biol. Cryst. Commun. 2005, 61, 981–984. [Google Scholar] [CrossRef]
  55. Drury, J.; Narayanaswami, V. Examination of lipid-bound conformation of apolipoprotein E4 by pyrene excimer fluorescence. J. Biol. Chem. 2005, 280, 14605–14610. [Google Scholar] [CrossRef]
  56. Krul, E.S.; Tikkanen, M.J.; Schonfeld, G. Heterogeneity of apolipoprotein E epitope expression on human lipoproteins: Importance for apolipoprotein E function. J. Lipid Res. 1988, 29, 1309–1325. [Google Scholar]
  57. Saito, H.; Dhanasekaran, P.; Baldwin, F.; Weisgraber, K.H.; Lund-Katz, S.; Phillips, M.C. Lipid binding-induced conformational change in human apolipoprotein E evidence for two lipid-bound states on spherical particles. J. Biol. Chem. 2001, 276, 40949–40954. [Google Scholar] [CrossRef] [PubMed]
  58. Raussens, V.; Drury, J.; Forte, T.M.; Choy, N.; Goormaghtigh, E.; Ruysschaert, J.-M.; Narayanaswami, V. Orientation and mode of lipid-binding interaction of human apolipoprotein E C-terminal domain. Biochem. J. 2005, 387, 747–754. [Google Scholar] [CrossRef] [PubMed]
  59. Frieden, C.; Wang, H.; Ho, C.M.W. A mechanism for lipid binding to apoE and the role of intrinsically disordered regions coupled to domain–domain interactions. Proc. Natl. Acad. Sci. USA 2017, 114, 6292–6297. [Google Scholar] [CrossRef] [PubMed]
  60. Matsuura, F.; Wang, N.; Chen, W.; Jiang, X.-C.; Tall, A.R. HDL from CETP-deficient subjects shows enhanced ability to promote cholesterol efflux from macrophages in an apoE- and ABCG1-dependent pathway. J. Clin. Investig. 2006, 116, 1435–1442. [Google Scholar] [CrossRef] [PubMed]
  61. Mahley, R.W.; Huang, Y.; Weisgraber, K.H. Putting cholesterol in its place: apoE and reverse cholesterol transport. J. Clin. Investig. 2006, 116, 1226–1229. [Google Scholar] [CrossRef]
  62. Ang, L.S.; Cruz, R.P.; Hendel, A.; Granville, D.J. Apolipoprotein E, an important player in longevity and age-related diseases. Exp. Gerontol. 2008, 43, 615–622. [Google Scholar] [CrossRef]
  63. Ilaria, Z.; Matteo, P.; Francesco, P.; Grazia, S.; Monica, G.; Laura, C.; Franco, B. Macrophage, but not systemic, apolipoprotein E is necessary for macrophage reverse cholesterol transport in vivo. Arterioscler. Thromb. Vasc. Biol. 2011, 31, 74–80. [Google Scholar]
  64. Spinney, L. Alzheimer’s disease: The forgetting gene. Nat. News 2014, 510, 26. [Google Scholar] [CrossRef]
  65. Raber, J.; Huang, Y.; Ashford, J.W. ApoE genotype accounts for the vast majority of AD risk and AD pathology. Neurobiol. Aging 2004, 25, 641–650. [Google Scholar] [CrossRef] [PubMed]
  66. Saunders, A.M.; Strittmatter, W.J.; Schmechel, D.; George-Hyslop, P.H.; Pericak-Vance, M.A.; Joo, S.H.; Rosi, B.L.; Gusella, J.F.; Crapper-MacLachlan, D.R.; Alberts, M.J. Association of apolipoprotein E allele ε4 with late-onset familial and sporadic Alzheimer’s disease. Neurology 1993, 43, 1467–1472. [Google Scholar] [CrossRef]
  67. Hyman, B.T.; Gomez-Isla, T.; West, H.; Briggs, M.; Chung, H.; Growdon, J.H.; Rebeck, G.W. Clinical and neuropathological correlates of apolipoprotein E genotype in Alzheimer’s disease. Window on molecular epidemiology. Ann. N. Y. Acad. Sci. 1996, 777, 158–165. [Google Scholar] [CrossRef]
  68. van Duijn, C.M.; de Knijff, P.; Cruts, M.; Wehnert, A.; Havekes, L.M.; Hofman, A.; Broeckhoven, C.V. Apolipoprotein E4 allele in a population-based study of early-onset Alzheimer’s disease. Nat. Genet. 1994, 7, 74. [Google Scholar] [CrossRef]
  69. Piedrahita, J.A.; Zhang, S.H.; Hagaman, J.R.; Oliver, P.M.; Maeda, N. Generation of mice carrying a mutant apolipoprotein E gene inactivated by gene targeting in embryonic stem cells. Proc. Natl. Acad. Sci. USA 1992, 89, 4471–4475. [Google Scholar] [CrossRef] [PubMed]
  70. Zhang, S.H.; Reddick, R.L.; Piedrahita, J.A.; Maeda, N. Spontaneous hypercholesterolemia and arterial lesions in mice lacking apolipoprotein E. Science 1992, 258, 468–471. [Google Scholar] [CrossRef] [PubMed]
  71. Hayek, T.; Oiknine, J.; Brook, J.G.; Aviram, M. Increased plasma and lipoprotein lipid peroxidation in apo E-deficient mice. Biochem. Biophys. Res. Commun. 1994, 201, 1567–1574. [Google Scholar] [CrossRef] [PubMed]
  72. Roselaar, S.E.; Daugherty, A. Apolipoprotein E-deficient mice have impaired innate immune responses to Listeria monocytogenes in vivo. J. Lipid Res. 1998, 39, 1740–1743. [Google Scholar] [PubMed]
  73. de Bont, N.; Netea, M.G.; Demacker, P.N.M.; Verschueren, I.; Kullberg, B.J.; van Dijk, K.W.; van der Meer, J.W.M.; Stalenhoef, A.F.H. Apolipoprotein E knock-out mice are highly susceptible to endotoxemia and Klebsiella pneumoniae infection. J. Lipid Res. 1999, 40, 680–685. [Google Scholar] [CrossRef]
  74. Robertson, T.A.; Dutton, N.S.; Martins, R.N.; Taddei, K.; Papadimitriou, J.M. Comparison of astrocytic and myocytic metabolic dysregulation in apolipoprotein E deficient and human apolipoprotein E transgenic mice. Neuroscience 2000, 98, 353–359. [Google Scholar] [CrossRef]
  75. Moghadasian, M.H.; Mcmanus, B.M.; Nguyen, L.B.; Shefer, S.; Nadji, M.; Godin, D.V.; Green, T.J.; Hill, J.; Yang, Y.; Scudamore, C.H.; et al. Pathophysiology of apolipoprotein E deficiency in mice: Relevance to apo E-related disorders in humans. FASEB J. 2001, 15, 2623–2630. [Google Scholar] [CrossRef] [PubMed]
  76. Kulminski, A.M.; Loika, Y.; Culminskaya, I.; Huang, J.; Arbeev, K.G.; Bagley, O.; Feitosa, M.F.; Zmuda, J.M.; Christensen, K.; Yashin, A.I. Independent associations of TOMM40 and APOE variants with body mass index. Aging Cell 2019, 18, e12869. [Google Scholar] [CrossRef]
  77. Harman, D. Aging: Overview. Ann. N. Y. Acad. Sci. 2001, 928, 1–21. [Google Scholar] [CrossRef]
  78. Vasto, S.; Candore, G.; Balistreri, C.R.; Caruso, M.; Colonna-Romano, G.; Grimaldi, M.P.; Listi, F.; Nuzzo, D.; Lio, D.; Caruso, C. Inflammatory networks in ageing, age-related diseases and longevity. Mech. Ageing Dev. 2007, 128, 83–91. [Google Scholar] [CrossRef]
  79. Kregel, K.C.; Zhang, H.J. An integrated view of oxidative stress in aging: Basic mechanisms, functional effects, and pathological considerations. Am. J. Physiol.-Regul. Integr. Comp. Physiol. 2007, 292, R18–R36. [Google Scholar] [CrossRef]
  80. Broer, L.; Buchman, A.S.; Deelen, J.; Evans, D.S.; Faul, J.D.; Lunetta, K.L.; Sebastiani, P.; Smith, J.A.; Smith, A.V.; Tanaka, T.; et al. GWAS of longevity in CHARGE consortium confirms APOE and FOXO3 candidacy. J. Gerontol. A. Biol. Sci. Med. Sci. 2015, 70, 110–118. [Google Scholar] [CrossRef] [PubMed]
  81. Iurescia, S.; Fioretti, D.; Mangialasche, F.; Rinaldi, M. The pathological cross talk between apolipoprotein E and amyloid-β peptide in Alzheimer’s disease: Emerging gene-based therapeutic approaches. J. Alzheimers Dis. 2010, 21, 35–48. [Google Scholar] [CrossRef] [PubMed]
  82. Dafnis, I.; Stratikos, E.; Tzinia, A.; Tsilibary, E.C.; Zannis, V.I.; Chroni, A. An apolipoprotein E4 fragment can promote intracellular accumulation of amyloid peptide β42. J. Neurochem. 2010, 115, 873–884. [Google Scholar] [CrossRef] [PubMed]
  83. Argyri, L.; Dafnis, I.; Theodossiou, T.A.; Gantz, D.; Stratikos, E.; Chroni, A. Molecular basis for increased risk for late-onset Alzheimer disease due to the naturally occurring L28P mutation in apolipoprotein E4. J. Biol. Chem. 2014, 289, 12931–12945. [Google Scholar] [CrossRef]
  84. Dafnis, I.; Argyri, L.; Sagnou, M.; Tzinia, A.; Tsilibary, E.C.; Stratikos, E.; Chroni, A. The ability of apolipoprotein E fragments to promote intraneuronal accumulation of amyloid β peptide 42 is both isoform and size-specific. Sci. Rep. 2016, 6, 30654. [Google Scholar] [CrossRef]
  85. Ji, Z.-S.; Miranda, R.D.; Newhouse, Y.M.; Weisgraber, K.H.; Huang, Y.; Mahley, R.W. Apolipoprotein E4 potentiates amyloid β peptide-induced lysosomal leakage and apoptosis in neuronal cells. J. Biol. Chem. 2002, 277, 21821–21828. [Google Scholar] [CrossRef]
  86. Ji, Z.-S.; Müllendorff, K.; Cheng, I.H.; Miranda, R.D.; Huang, Y.; Mahley, R.W. Reactivity of apolipoprotein E4 and amyloid β peptide lysosomal stability and neurodegeneration. J. Biol. Chem. 2006, 281, 2683–2692. [Google Scholar] [CrossRef]
  87. Buttini, M.; Orth, M.; Bellosta, S.; Akeefe, H.; Pitas, R.E.; Wyss-Coray, T.; Mucke, L.; Mahley, R.W. Expression of human apolipoprotein E3 or E4 in the brains ofApoE−/− mice: Isoform-specific effects on neurodegeneration. J. Neurosci. 1999, 19, 4867–4880. [Google Scholar] [CrossRef] [PubMed]
  88. Hartman, R.E.; Wozniak, D.F.; Nardi, A.; Olney, J.W.; Sartorius, L.; Holtzman, D.M. Behavioral phenotyping of GFAP-ApoE3 and -ApoE4 transgenic mice: ApoE4 mice show profound working memory impairments in the absence of Alzheimer’s-like neuropathology. Exp. Neurol. 2001, 170, 326–344. [Google Scholar] [CrossRef] [PubMed]
  89. Bour, A.; Grootendorst, J.; Vogel, E.; Kelche, C.; Dodart, J.-C.; Bales, K.; Moreau, P.-H.; Sullivan, P.M.; Mathis, C. Middle-aged human apoE4 targeted-replacement mice show retention deficits on a wide range of spatial memory tasks. Behav. Brain Res. 2008, 193, 174–182. [Google Scholar] [CrossRef] [PubMed]
  90. Raber, J.; Wong, D.; Buttini, M.; Orth, M.; Bellosta, S.; Pitas, R.E.; Mahley, R.W.; Mucke, L. Isoform-specific effects of human apolipoprotein E on brain function revealed in ApoE knockout mice: Increased susceptibility of females. Proc. Natl. Acad. Sci. USA 1998, 95, 10914–10919. [Google Scholar] [CrossRef]
  91. Nathan, B.P.; Bellosta, S.; Sanan, D.A.; Weisgraber, K.H.; Mahley, R.W.; Pitas, R.E. Differential effects of apolipoproteins E3 and E4 on neuronal growth in vitro. Science 1994, 264, 850–852. [Google Scholar] [CrossRef] [PubMed]
  92. Nathan, B.P.; Chang, K.-C.; Bellosta, S.; Brisch, E.; Ge, N.; Mahley, R.W.; Pitas, R.E. The inhibitory effect of apolipoprotein E4 on neurite outgrowth is associated with microtubule depolymerization. J. Biol. Chem. 1995, 270, 19791–19799. [Google Scholar] [CrossRef] [PubMed]
  93. Holtzman, D.M.; Pitas, R.E.; Kilbridge, J.; Nathan, B.; Mahley, R.W.; Bu, G.; Schwartz, A.L. Low density lipoprotein receptor-related protein mediates apolipoprotein E-dependent neurite outgrowth in a central nervous system-derived neuronal cell line. Proc. Natl. Acad. Sci. USA 1995, 92, 9480–9484. [Google Scholar] [CrossRef]
  94. Li, G.; Bien-Ly, N.; Andrews-Zwilling, Y.; Xu, Q.; Bernardo, A.; Ring, K.; Halabisky, B.; Deng, C.; Mahley, R.W.; Huang, Y. GABAergic interneuron dysfunction impairs hippocampal neurogenesis in adult apolipoprotein E4 knockin mice. Cell Stem Cell 2009, 5, 634–645. [Google Scholar] [CrossRef] [PubMed]
  95. Reiman, E.M.; Chen, K.; Alexander, G.E.; Caselli, R.J.; Bandy, D.; Osborne, D.; Saunders, A.M.; Hardy, J. Correlations between apolipoprotein E ε4 gene dose and brain-imaging measurements of regional hypometabolism. Proc. Natl. Acad. Sci. USA 2005, 102, 8299–8302. [Google Scholar] [CrossRef]
  96. Chang, S.; Ran Ma, T.; Miranda, R.D.; Balestra, M.E.; Mahley, R.W.; Huang, Y. Lipid- and receptor-binding regions of apolipoprotein E4 fragments act in concert to cause mitochondrial dysfunction and neurotoxicity. Proc. Natl. Acad. Sci. USA 2005, 102, 18694–18699. [Google Scholar] [CrossRef]
  97. Scarmeas, N.; Habeck, C.; Hilton, J.; Anderson, K.; Flynn, J.; Park, A.; Stern, Y. APOE related alterations in cerebral activation even at college age. J. Neurol. Neurosurg. Psychiatry 2005, 76, 1440–1444. [Google Scholar] [CrossRef]
  98. Nakamura, T.; Watanabe, A.; Fujino, T.; Hosono, T.; Michikawa, M. Apolipoprotein E4 (1–272) fragment is associated with mitochondrial proteins and affects mitochondrial function in neuronal cells. Mol. Neurodegener. 2009, 4, 35. [Google Scholar] [CrossRef]
  99. Tanaka, M.; Vedhachalam, C.; Sakamoto, T.; Dhanasekaran, P.; Phillips, M.C.; Lund-Katz, S.; Saito, H. Effect of carboxyl-terminal truncation on structure and lipid interaction of human apolipoprotein E4. Biochemistry 2006, 45, 4240–4247. [Google Scholar] [CrossRef]
  100. Chou, C.-Y.; Jen, W.-P.; Hsieh, Y.-H.; Shiao, M.-S.; Chang, G.-G. Structural and functional variations in human apolipoprotein E3 and E4. J. Biol. Chem. 2006, 281, 13333–13344. [Google Scholar] [CrossRef]
  101. Tambini, M.D.; Pera, M.; Kanter, E.; Yang, H.; Guardia-Laguarta, C.; Holtzman, D.; Sulzer, D.; Area-Gomez, E.; Schon, E.A. ApoE4 upregulates the activity of mitochondria-associated ER membranes. EMBO Rep. 2016, 17, 27–36. [Google Scholar] [CrossRef] [PubMed]
  102. Bravo, R.; Gutierrez, T.; Paredes, F.; Gatica, D.; Rodriguez, A.E.; Pedrozo, Z.; Chiong, M.; Parra, V.; Quest, A.F.G.; Rothermel, B.A.; et al. Endoplasmic reticulum: ER stress regulates mitochondrial bioenergetics. Int. J. Biochem. Cell Biol. 2012, 44, 16–20. [Google Scholar] [CrossRef] [PubMed]
  103. Cao, S.S.; Kaufman, R.J. Endoplasmic reticulum stress and oxidative stress in cell fate decision and human disease. Antioxid. Redox Signal. 2014, 21, 396–413. [Google Scholar] [CrossRef] [PubMed]
  104. Chaudhari, N.; Talwar, P.; Parimisetty, A.; Lefebvre d’Hellencourt, C.; Ravanan, P. A Molecular web: Endoplasmic reticulum stress, inflammation, and oxidative stress. Front. Cell. Neurosci. 2014, 8. [Google Scholar] [CrossRef]
  105. Miyata, M.; Smith, J.D. Apolipoprotein E allele–specific antioxidant activity and effects on cytotoxicity by oxidative insults and β–amyloid peptides. Nat. Genet. 1996, 14, 55–61. [Google Scholar] [CrossRef] [PubMed]
  106. Jofre-Monseny, L.; Loboda, A.; Wagner, A.E.; Huebbe, P.; Boesch-Saadatmandi, C.; Jozkowicz, A.; Minihane, A.-M.; Dulak, J.; Rimbach, G. Effects of apoE genotype on macrophage inflammation and heme oxygenase-1 expression. Biochem. Biophys. Res. Commun. 2007, 357, 319–324. [Google Scholar] [CrossRef]
  107. Itzhaki, R.F. Herpes and Alzheimer’s disease: Subversion in the central nervous system and how it might be halted. J. Alzheimers Dis. JAD 2016, 54, 1273–1281. [Google Scholar] [CrossRef]
  108. Itzhaki, R.F. Herpes simplex virus type 1 and Alzheimer’s disease: Possible mechanisms and signposts. FASEB J. 2017, 31, 3216–3226. [Google Scholar] [CrossRef] [PubMed]
  109. Itzhaki, R.F. Corroboration of a major role for Herpes simplex virus type 1 in Alzheimer’s disease. Front. Aging Neurosci. 2018, 10, 324. [Google Scholar] [CrossRef] [PubMed]
  110. Cun, W.; Jiang, J.; Luo, G. The C-terminal α-helix domain of apolipoprotein E Is required for interaction with nonstructural protein 5A and assembly of Hepatitis C virus. J. Virol. 2010, 84, 11532–11541. [Google Scholar] [CrossRef] [PubMed]
  111. Chiba-Falek, O.; Linnertz, C.; Guyton, J.; Gardner, S.D.; Roses, A.D.; McCarthy, J.J.; Patel, K. Pleiotropy and allelic heterogeneity in the TOMM40-APOE genomic region related to clinical and metabolic features of hepatitis C infection. Hum. Genet. 2012, 131, 1911–1920. [Google Scholar] [CrossRef]
  112. Bankwitz, D.; Doepke, M.; Hueging, K.; Weller, R.; Bruening, J.; Behrendt, P.; Lee, J.-Y.; Vondran, F.W.R.; Manns, M.P.; Bartenschlager, R.; et al. Maturation of secreted HCV particles by incorporation of secreted ApoE protects from antibodies by enhancing infectivity. J. Hepatol. 2017, 67, 480–489. [Google Scholar] [CrossRef]
  113. Weller, R.; Hueging, K.; Brown, R.J.P.; Todt, D.; Joecks, S.; Vondran, F.W.R.; Pietschmann, T. Hepatitis C virus strain-dependent usage of apolipoprotein E modulates assembly efficiency and specific infectivity of secreted virions. J. Virol. 2017, 91. [Google Scholar] [CrossRef]
  114. Gondar, V.; Molina-Jiménez, F.; Hishiki, T.; García-Buey, L.; Koutsoudakis, G.; Shimotohno, K.; Benedicto, I.; Majano, P.L. Apolipoprotein E, but not apolipoprotein B, Is essential for efficient cell-to-cell transmission of Hepatitis C virus. J. Virol. 2015, 89, 9962–9973. [Google Scholar] [CrossRef]
  115. Popescu, C.-I.; Dubuisson, J. Role of lipid metabolism in hepatitis C virus assembly and entry. Biol. Cell 2010, 102, 63–74. [Google Scholar] [CrossRef] [PubMed]
  116. Chang, L.; Andres, M.; Sadino, J.; Jiang, C.S.; Nakama, H.; Miller, E.; Ernst, T. Impact of apolipoprotein E ε4 and HIV on cognition and brain atrophy: antagonistic pleiotropy and premature brain aging. NeuroImage 2011, 58, 1017–1027. [Google Scholar] [CrossRef]
  117. Chang, L.; Jiang, C.; Cunningham, E.; Buchthal, S.; Douet, V.; Andres, M.; Ernst, T. Effects of APOE ε4, age, and HIV on glial metabolites and cognitive deficits. Neurology 2014, 82, 2213–2222. [Google Scholar] [CrossRef]
  118. Wendelken, L.A.; Jahanshad, N.; Rosen, H.J.; Busovaca, E.; Allen, I.; Coppola, G.; Adams, C.; Rankin, K.P.; Milanini, B.; Clifford, K.; et al. ApoE ε4 is associated with cognition, brain integrity and atrophy in HIV over age 60. J. Acquir. Immune Defic. Syndr. 1999 2016, 73, 426–432. [Google Scholar] [CrossRef]
  119. Suwalak, T.; Srisawasdi, P.; Puangpetch, A.; Santon, S.; Koomdee, N.; Chamnanphon, M.; Charoenyingwattana, A.; Chantratita, W.; Sukasem, C. Polymorphisms of the ApoE (Apolipoprotein E) gene and their influence on dyslipidemia in hiv-1-infected individuals. Jpn. J. Infect. Dis. 2015, 68, 5–12. [Google Scholar] [CrossRef]
  120. Cooley, S.A.; Paul, R.H.; Fennema-Notestine, C.; Morgan, E.E.; Vaida, F.; Deng, Q.; Chen, J.A.; Letendre, S.; Ellis, R.; Clifford, D.B.; et al. Apolipoprotein E ε4 genotype status is not associated with neuroimaging outcomes in a large cohort of HIV+ individuals. J. Neurovirol. 2016, 22, 607–614. [Google Scholar] [CrossRef] [PubMed]
  121. Geffin, R.; McCarthy, M. Aging and apolipoprotein E in HIV infection. J. Neurovirol. 2018, 24, 529–548. [Google Scholar] [CrossRef] [PubMed]
  122. Stengard, J.H.; Weiss, K.M.; Sing, C.F. An ecological study of association between coronary heart disease mortality rates in men and the relative frequencies of common allelic variations in the gene coding for apolipoprotein E. Hum. Genet. 1998, 103, 234–241. [Google Scholar] [CrossRef]
  123. Gerdes, L.U.; Gerdes, C.; Kervinen, K.; Savolainen, M.; Klausen, I.C.; Hansen, P.S.; Kesäniemi, Y.A.; Faergeman, O. The apolipoprotein ε4 allele determines prognosis and the effect on prognosis of simvastatin in survivors of myocardial infarction: A substudy of the Scandinavian simvastatin survival study. Circulation 2000, 101, 1366–1371. [Google Scholar] [CrossRef] [PubMed]
  124. Kumar, N.T.; Liestøl, K.; Løberg, E.M.; Reims, H.M.; Brorson, S.-H.; Mæhlen, J. The apolipoprotein E polymorphism and cardiovascular diseases—An autopsy study. Cardiovasc. Pathol. 2012, 21, 461–469. [Google Scholar] [CrossRef] [PubMed]
  125. Satizabal Claudia, L.; Samieri, C.; Davis-Plourde, K.L.; Voetsch, B.; Aparicio, H.J.; Pase, M.P.; Romero, J.R.; Helmer, C.; Vasan, R.S.; Kase, C.S.; et al. APOE and the association of fatty acids with the risk of stroke, coronary heart disease, and mortality. Stroke 2018, 49, 2822–2829. [Google Scholar] [CrossRef] [PubMed]
  126. Foraker, J.; Millard, S.P.; Leong, L.; Thomson, Z.; Chen, S.; Keene, C.D.; Bekris, L.M.; Yu, C.-E. The APOE gene is differentially methylated in Alzheimer’s disease. J. Alzheimers Dis. 2015, 48, 745–755. [Google Scholar] [CrossRef]
  127. Liu, J.; Zhao, W.; Ware, E.B.; Turner, S.T.; Mosley, T.H.; Smith, J.A. DNA methylation in the APOE genomic region is associated with cognitive function in African Americans. BMC Med. Genomics 2018, 11. [Google Scholar] [CrossRef]
  128. Giuliani, C.; Sazzini, M.; Pirazzini, C.; Bacalini, M.G.; Marasco, E.; Ruscone, G.A.G.; Fang, F.; Sarno, S.; Gentilini, D.; Di Blasio, A.M.; et al. Impact of demography and population dynamics on the genetic architecture of human longevity. Aging 2018, 10, 1947–1963. [Google Scholar] [CrossRef]
  129. Sebastiani, P.; Solovieff, N.; DeWan, A.T.; Walsh, K.M.; Puca, A.; Hartley, S.W.; Melista, E.; Andersen, S.; Dworkis, D.A.; Wilk, J.B.; et al. Genetic Signatures of exceptional longevity in humans. PLoS ONE 2012, 7. [Google Scholar] [CrossRef]
  130. Revelas, M.; Thalamuthu, A.; Oldmeadow, C.; Evans, T.-J.; Armstrong, N.J.; Kwok, J.B.; Brodaty, H.; Schofield, P.R.; Scott, R.J.; Sachdev, P.S.; et al. Review and meta-analysis of genetic polymorphisms associated with exceptional human longevity. Mech. Ageing Dev. 2018, 175, 24–34. [Google Scholar] [CrossRef]
  131. Pilling, L.C.; Atkins, J.L.; Bowman, K.; Jones, S.E.; Tyrrell, J.; Beaumont, R.N.; Ruth, K.S.; Tuke, M.A.; Yaghootkar, H.; Wood, A.R.; et al. Human longevity is influenced by many genetic variants: Evidence from 75,000 UK Biobank participants. Aging 2016, 8, 547–560. [Google Scholar] [CrossRef]
  132. Pilling, L.C.; Kuo, C.-L.; Sicinski, K.; Tamosauskaite, J.; Kuchel, G.A.; Harries, L.W.; Herd, P.; Wallace, R.; Ferrucci, L.; Melzer, D. Human longevity: 25 genetic loci associated in 389,166 UK biobank participants. Aging 2017, 9, 2504–2520. [Google Scholar] [CrossRef]
  133. Newman, A.B.; Walter, S.; Lunetta, K.L.; Garcia, M.E.; Slagboom, P.E.; Christensen, K.; Arnold, A.M.; Aspelund, T.; Aulchenko, Y.S.; Benjamin, E.J.; et al. 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. Sci. 2010, 65A, 478–487. [Google Scholar] [CrossRef] [PubMed]
  134. De Benedictis, G.; Franceschi, C. The unusual genetics of human longevity. Sci. Aging Knowl. Environ. 2006, 2006, pe20. [Google Scholar] [CrossRef]
  135. Zeng, Y.; Nie, C.; Min, J.; Liu, X.; Li, M.; Chen, H.; Xu, H.; Wang, M.; Ni, T.; Li, Y.; et al. Novel loci and pathways significantly associated with longevity. Sci. Rep. 2016, 6, 21243. [Google Scholar] [CrossRef] [PubMed]
  136. Giuliani, C.; Garagnani, P.; Franceschi, C. Genetics of Human longevity within an eco-evolutionary nature-nurture framework. Circ. Res. 2018, 123, 745–772. [Google Scholar] [CrossRef] [PubMed]
  137. Fuku, N.; Díaz-Peña, R.; Arai, Y.; Abe, Y.; Zempo, H.; Naito, H.; Murakami, H.; Miyachi, M.; Spuch, C.; Serra-Rexach, J.A.; et al. Epistasis, physical capacity-related genes and exceptional longevity: FNDC5 gene interactions with candidate genes FOXOA3 and APOE. BMC Genomics 2017, 18, 803. [Google Scholar] [CrossRef]
  138. Fortney, K.; Dobriban, E.; Garagnani, P.; Pirazzini, C.; Monti, D.; Mari, D.; Atzmon, G.; Barzilai, N.; Franceschi, C.; Owen, A.B.; et al. Genome-wide scan informed by age-related disease identifies loci for exceptional human longevity. PLoS Genet. 2015, 11. [Google Scholar] [CrossRef] [PubMed]
  139. Wang, J.; Shi, L.; Zou, Y.; Tang, J.; Cai, J.; Wei, Y.; Qin, J.; Zhang, Z. Positive association of familial longevity with the moderate-high HDL-C concentration in Bama aging Study. Aging 2018, 10, 3528–3540. [Google Scholar] [CrossRef]
  140. Silva-Sena, G.G.; Camporez, D.; dos Santos, L.R.; da Silva, A.S.; Sagrillo Pimassoni, L.H.; Tieppo, A.; do Pimentel Batitucci, M.C.; Morelato, R.L.; de Paula, F. An association study of FOXO3 variant and longevity. Genet. Mol. Biol. 2018, 41, 386–396. [Google Scholar] [CrossRef]
  141. Garatachea, N.; Emanuele, E.; Calero, M.; Fuku, N.; Arai, Y.; Abe, Y.; Murakami, H.; Miyachi, M.; Yvert, T.; Verde, Z.; et al. ApoE gene and exceptional longevity: Insights from three independent cohorts. Exp. Gerontol. 2014, 53, 16–23. [Google Scholar] [CrossRef]
  142. Ryu, S.; Atzmon, G.; Barzilai, N.; Raghavachari, N.; Suh, Y. Genetic landscape of APOE in human longevity revealed by high-throughput sequencing. Mech. Ageing Dev. 2016, 155, 7–9. [Google Scholar] [CrossRef]
  143. Louhija, J.; Miettinen, H.E.; Kontula, K.; Tikkanen, M.J.; Miettinen, T.A.; Tilvis, R.S. Aging and genetic variation of plasma apolipoproteins. Relative loss of the apolipoprotein E4 phenotype in centenarians. Arterioscler. Thromb. J. Vasc. Biol. 1994, 14, 1084–1089. [Google Scholar] [CrossRef]
  144. Schächter, F.; Faure-Delanef, L.; Guénot, F.; Rouger, H.; Froguel, P.; Lesueur-Ginot, L.; Cohen, D. Genetic associations with human longevity at the APOE and ACE loci. Nat. Genet. 1994, 6, 29–32. [Google Scholar] [CrossRef] [PubMed]
  145. Asada, T.; Kariya, T.; Yamagata, Z.; Kinoshita, T.; Asaka, A. Apolipoprotein E allele in centenarians. Neurology 1996, 46, 1484. [Google Scholar] [CrossRef] [PubMed]
  146. Yamagata, Z.; Asada, T.; Kinoshita, A.; Zhang, Y.; Asaka, A. Distribution of apolipoprotein E gene polymorphisms in Japanese patients with Alzheimer’s disease and in Japanese centenarians. Hum. Hered. 1997, 47, 22–26. [Google Scholar] [CrossRef] [PubMed]
  147. Castro, E.; Ogburn, C.E.; Hunt, K.E.; Tilvis, R.; Louhija, J.; Penttinen, R.; Erkkola, R.; Panduro, A.; Riestra, R.; Piussan, C.; et al. Polymorphisms at the Werner locus: I. Newly identified polymorphisms, ethnic variability of 1367Cy/Arg, and its stability in a population of Finnish centenarians. Am. J. Med. Genet. 1999, 82, 399–403. [Google Scholar] [CrossRef]
  148. Gerdes, L.U.; Jeune, B.; Ranberg, K.A.; Nybo, H.; Vaupel, J.W. 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. Epidemiol. 2000, 19, 202–210. [Google Scholar] [CrossRef]
  149. Blanché, H.; Cabanne, L.; Sahbatou, M.; Thomas, G. A study of French centenarians: Are ACE and APOE associated with longevity? Comptes Rendus Académie Sci. Ser. III Sci. Vie 2001, 324, 129–135. [Google Scholar] [CrossRef]
  150. Arai, Y.; Hirose, N.; Nakazawa, S.; Yamamura, K.; Shimizu, K.; Takayama, M.; Ebihara, Y.; Osono, Y.; Homma, S. Lipoprotein metabolism in Japanese centenarians: Effects of apolipoprotein E polymorphism and nutritional status. J. Am. Geriatr. Soc. 2001, 49, 1434–1441. [Google Scholar] [CrossRef] [PubMed]
  151. Choi, Y.-H.; Kim, J.-H.; Kim, D.K.; Kim, J.-W.; Kim, D.-K.; Lee, M.S.; Kim, C.H.; Park, S.C. Distributions of ACE and APOE polymorphisms and their relations with dementia status in Korean centenarians. J. Gerontol. Ser. A 2003, 58, M227–M231. [Google Scholar] [CrossRef]
  152. Panza, F.; Solfrizzi, V.; Colacicco, A.M.; Basile, A.M.; D’Introno, A.; Capurso, C.; Sabba, M.; Capurso, S.; Capurso, A. Apolipoprotein E (APOE) polymorphism influences serum APOE levels in Alzheimer’s disease patients and centenarians. NeuroReport 2003, 14, 605. [Google Scholar] [CrossRef]
  153. Capurso, C.; Solfrizzi, V.; D’Introno, A.; Colacicco, A.M.; Capurso, S.A.; Semeraro, C.; Capurso, A.; Panza, F. Interleukin 6−174 G/C promoter gene polymorphism in centenarians: No evidence of association with human longevity or interaction with apolipoprotein E alleles. Exp. Gerontol. 2004, 39, 1109–1114. [Google Scholar] [CrossRef]
  154. Garatachea, N.; Marín, P.J.; 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. 2014, 18, 3–13. [Google Scholar] [CrossRef]
  155. Rea, I.M.; Mc Dowell, I.; McMaster, D.; Smye, M.; Stout, R.; Evans, A. Apolipoprotein E alleles in nonagenarian subjects in the Belfast Elderly Longitudinal Free-living Ageing Study (BELFAST). Mech. Ageing Dev. 2001, 122, 1367–1372. [Google Scholar] [CrossRef]
  156. Zubenko, G.S.; Stiffler, J.S.; Hughes, H.B.; Fatigati, M.J.; Zubenko, W.N. Genome survey for loci that influence successful aging: sample characterization, method validation, and initial results for the Y chromosome. Am. J. Geriatr. Psychiatry 2002, 10, 619–630. [Google Scholar] [CrossRef]
  157. Geesaman, B.J.; Benson, E.; Brewster, S.J.; Kunkel, L.M.; Blanché, H.; Thomas, G.; Perls, T.T.; Daly, M.J.; Puca, A.A. Haplotype-based identification of a microsomal transfer protein marker associated with the human lifespan. Proc. Natl. Acad. Sci. USA 2003, 100, 14115–14120. [Google Scholar] [CrossRef] [PubMed]
  158. Feng, J.; Xiang, L.; Wan, G.; Qi, K.; Sun, L.; Huang, Z.; Zheng, C.; Lv, Z.; Hu, C.; Yang, Z. Is APOE ε3 a favourable factor for the longevity: An association study in Chinese population. J. Genet. 2011, 90, 343–347. [Google Scholar] [CrossRef] [PubMed]
  159. Sebastiani, P.; Gurinovich, A.; Nygaard, M.; Sasaki, T.; Sweigart, B.; Bae, H.; Andersen, S.L.; Villa, F.; Atzmon, G.; Christensen, K.; et al. APOE alleles and extreme human longevity. J. Gerontol. Ser. A 2019, 74, 44–51. [Google Scholar] [CrossRef]
  160. Mostafavi, H.; Berisa, T.; Day, F.R.; Perry, J.R.B.; Przeworski, M.; Pickrell, J.K. Identifying genetic variants that affect viability in large cohorts. PLoS Biol. 2017, 15, e2002458. [Google Scholar] [CrossRef] [PubMed]
  161. Christensen, K.; Johnson, T.E.; Vaupel, J.W. The quest for genetic determinants of human longevity: Challenges and insights. Nat. Rev. Genet. 2006, 7, 436–448. [Google Scholar] [CrossRef]
  162. Huebbe, P.; Rimbach, G. Evolution of human apolipoprotein E (APOE) isoforms: Gene structure, protein function and interaction with dietary factors. Ageing Res. Rev. 2017, 37, 146–161. [Google Scholar] [CrossRef]
  163. Luo, C.-C.; Li, W.-H.; Moore, M.N.; Chan, L. Structure and evolution of the apolipoprotein multigene family. J. Mol. Biol. 1986, 187, 325–340. [Google Scholar] [CrossRef]
  164. Smith, A.F.; Owen, L.M.; Strobel, L.M.; Chen, H.; Kanost, M.R.; Hanneman, E.; Wells, M.A. Exchangeable apolipoproteins of insects share a common structural motif. J. Lipid Res. 1994, 35, 1976–1984. [Google Scholar]
  165. Peterson, K.J.; Lyons, J.B.; Nowak, K.S.; Takacs, C.M.; Wargo, M.J.; McPeek, M.A. Estimating metazoan divergence times with a molecular clock. Proc. Natl. Acad. Sci. USA 2004, 101, 6536–6541. [Google Scholar] [CrossRef]
  166. Reich, D.; Green, R.E.; Kircher, M.; Krause, J.; Patterson, N.; Durand, E.Y.; Viola, B.; Briggs, A.W.; Stenzel, U.; Johnson, P.L.F.; et al. Genetic history of an archaic hominin group from Denisova Cave in Siberia. Nature 2010, 468, 1053–1060. [Google Scholar] [CrossRef]
  167. McIntosh, A.M.; Bennett, C.; Dickson, D.; Anestis, S.F.; Watts, D.P.; Webster, T.H.; Fontenot, M.B.; Bradley, B.J. The apolipoprotein E (APOE) gene appears functionally monomorphic in chimpanzees (Pan troglodytes). PLoS ONE 2012, 7. [Google Scholar] [CrossRef]
  168. Finch, C.E.; Stanford, C.B. Meat-Adaptive genes and the evolution of slower aging in humans. Q. Rev. Biol. 2004, 79, 3–50. [Google Scholar] [CrossRef]
  169. Raichlen, D.A.; Alexander, G.E. Exercise, APOE genotype, and the evolution of the human lifespan. Trends Neurosci. 2014, 37, 247–255. [Google Scholar] [CrossRef]
  170. Fullerton, S.M.; Clark, A.G.; Weiss, K.M.; Nickerson, D.A.; Taylor, S.L.; Stengård, J.H.; Salomaa, V.; Vartiainen, E.; Perola, M.; Boerwinkle, E.; et al. Apolipoprotein E variation at the sequence haplotype level: Implications for the origin and maintenance of a major human polymorphism. Am. J. Hum. Genet. 2000, 67, 881–900. [Google Scholar] [CrossRef]
  171. Antón, S.C.; Leonard, W.R.; Robertson, M.L. An ecomorphological model of the initial hominid dispersal from Africa. J. Hum. Evol. 2002, 43, 773–785. [Google Scholar] [CrossRef]
  172. Bramble, D.M.; Lieberman, D.E. Endurance running and the evolution of Homo. Nature 2004, 432, 345–352. [Google Scholar] [CrossRef]
  173. Malina, R.M.; Little, B.B. Physical activity: The present in the context of the past. Am. J. Hum. Biol. Off. J. Hum. Biol. Counc. 2008, 20, 373–391. [Google Scholar] [CrossRef]
  174. Caspari, R.; Lee, S.-H. Older age becomes common late in human evolution. Proc. Natl. Acad. Sci. USA 2004, 101, 10895–10900. [Google Scholar] [CrossRef]
  175. Hawkes, K.; O’Connell, J.F.; Jones, N.G.B.; Alvarez, H.; Charnov, E.L. Grandmothering, menopause, and the evolution of human life histories. Proc. Natl. Acad. Sci. USA 1998, 95, 1336–1339. [Google Scholar] [CrossRef]
  176. Hawkes, K. Genomic evidence for the evolution of human postmenopausal longevity. Proc. Natl. Acad. Sci. USA 2016, 113, 17–18. [Google Scholar] [CrossRef]
  177. Ojeda-Granados, C.; Panduro, A.; Gonzalez-Aldaco, K.; Sepulveda-Villegas, M.; Rivera-Iñiguez, I.; Roman, S. Tailoring nutritional advice for Mexicans based on prevalence profiles of diet-related adaptive gene polymorphisms. J. Pers. Med. 2017, 7, 16. [Google Scholar] [CrossRef]
  178. Singh, P.P.; Singh, M.; Mastana, S.S. APOE distribution in world populations with new data from India and the UK. Ann. Hum. Biol. 2006, 33, 279–308. [Google Scholar] [CrossRef]
  179. Hu, P.; Qin, Y.H.; Jing, C.X.; Lu, L.; Hu, B.; Du, P.F. Does the geographical gradient of ApoE4 allele exist in China? A systemic comparison among multiple Chinese populations. Mol. Biol. Rep. 2011, 38, 489–494. [Google Scholar] [CrossRef]
  180. Zekraoui, L.; Lagarde, J.P.; Raisonnier, A.; Gérard, N.; Aouizérate, A.; Lucotte, G. High frequency of the apolipoprotein E *4 allele in African pygmies and most of the African populations in sub-Saharan Africa. Hum. Biol. 1997, 69, 575–581. [Google Scholar]
  181. Corbo, R.M.; Scacchi, R. Apolipoprotein E (APOE) allele distribution in the world. Is APOE * 4 a ‘thrifty’ allele? Ann. Hum. Genet. 1999, 63, 301–310. [Google Scholar] [CrossRef] [PubMed]
  182. Sazzini, M.; Gnecchi Ruscone, G.A.; Giuliani, C.; Sarno, S.; Quagliariello, A.; De Fanti, S.; Boattini, A.; Gentilini, D.; Fiorito, G.; Catanoso, M.; et al. Complex interplay between neutral and adaptive evolution shaped differential genomic background and disease susceptibility along the Italian peninsula. Sci. Rep. 2016, 6, 32513. [Google Scholar] [CrossRef] [PubMed]
  183. Boattini, A.; Martinez-Cruz, B.; Sarno, S.; Harmant, C.; Useli, A.; Sanz, P.; Yang-Yao, D.; Manry, J.; Ciani, G.; Luiselli, D.; et al. Uniparental markers in Italy reveal a sex-biased genetic structure and different historical strata. PLoS ONE 2013, 8, e65441. [Google Scholar] [CrossRef]
  184. Ye, K.; Gao, F.; Wang, D.; Bar-Yosef, O.; Keinan, A. Dietary adaptation of FADS genes in Europe varied across time and geography. Nat. Ecol. Evol. 2017, 1, 167. [Google Scholar] [CrossRef]
  185. Buckley, M.T.; Racimo, F.; Allentoft, M.E.; Jensen, M.K.; Jonsson, A.; Huang, H.; Hormozdiari, F.; Sikora, M.; Marnetto, D.; Eskin, E.; et al. Selection in Europeans on fatty acid desaturases associated with dietary changes. Mol. Biol. Evol. 2017, 34, 1307–1318. [Google Scholar] [CrossRef]
  186. Egert, S.; Rimbach, G.; Huebbe, P. ApoE genotype: From geographic distribution to function and responsiveness to dietary factors. Proc. Nutr. Soc. 2012, 71, 410–424. [Google Scholar] [CrossRef]
  187. Graeser, A.-C.; Boesch-Saadatmandi, C.; Lippmann, J.; Wagner, A.E.; Huebbe, P.; Storm, N.; Höppner, W.; Wiswedel, I.; Gardemann, A.; Minihane, A.M.; et al. Nrf2-dependent gene expression is affected by the proatherogenic apoE4 genotype—studies in targeted gene replacement mice. J. Mol. Med. 2011, 89, 1027–1035. [Google Scholar] [CrossRef] [PubMed]
  188. Huebbe, P.; Nebel, A.; Siegert, S.; Moehring, J.; Boesch-Saadatmandi, C.; Most, E.; Pallauf, J.; Egert, S.; Müller, M.J.; Schreiber, S.; et al. APOE ε4 is associated with higher vitamin D levels in targeted replacement mice and humans. FASEB J. 2011, 25, 3262–3270. [Google Scholar] [CrossRef]
  189. Azevedo, O.G.R.; Bolick, D.T.; Roche, J.K.; Pinkerton, R.F.; Lima, A.A.M.; Vitek, M.P.; Warren, C.A.; Oriá, R.B.; Guerrant, R.L. Apolipoprotein E Plays a key role against cryptosporidial infection in transgenic undernourished mice. PLoS ONE 2014, 9. [Google Scholar] [CrossRef]
  190. Huebbe, P.; Dose, J.; Schloesser, A.; Campbell, G.; Glüer, C.-C.; Gupta, Y.; Ibrahim, S.; Minihane, A.-M.; Baines, J.F.; Nebel, A.; et al. Apolipoprotein E (APOE) genotype regulates body weight and fatty acid utilization—Studies in gene-targeted replacement mice. Mol. Nutr. Food Res. 2015, 59, 334–343. [Google Scholar] [CrossRef]
  191. Huebbe, P.; Lange, J.; Lietz, G.; Rimbach, G. Dietary β-carotene and lutein metabolism is modulated by the APOE genotype. BioFactors 2016, 42, 388–396. [Google Scholar] [CrossRef]
  192. Luca, F.; Perry, G.H.; Di Rienzo, A. Evolutionary adaptations to dietary changes. Annu. Rev. Nutr. 2010, 30, 291–314. [Google Scholar] [CrossRef] [PubMed]
  193. Finch, C.E.; Sapolsky, R.M. The evolution of Alzheimer disease, the reproductive schedule, and apoE isoforms. Neurobiol. Aging 1999, 20, 407–428. [Google Scholar] [CrossRef]
  194. Schwarz, F.; Springer, S.A.; Altheide, T.K.; Varki, N.M.; Gagneux, P.; Varki, A. Human-specific derived alleles of CD33 and other genes protect against postreproductive cognitive decline. Proc. Natl. Acad. Sci. 2016, 113, 74–79. [Google Scholar] [CrossRef] [PubMed]
  195. van Exel, E.; Koopman, J.J.E.; van Bodegom, D.; Meij, J.J.; de Knijff, P.; Ziem, J.B.; Finch, C.E.; Westendorp, R.G.J. Effect of APOE ε4 allele on survival and fertility in an adverse environment. PLoS ONE 2017, 12, e0179497. [Google Scholar] [CrossRef]
  196. Asgari, N.; Akbari, M.T.; Zare, S.; Babamohammadi, G. Positive association of apolipoprotein E4 polymorphism with recurrent pregnancy loss in Iranian patients. J. Assist. Reprod. Genet. 2013, 30, 265–268. [Google Scholar] [CrossRef] [PubMed]
  197. Goodman, C.; Goodman, C.S.; Hur, J.; Jeyendran, R.S.; Coulam, C. The Association of apoprotien E polymorphisms with recurrent pregnancy loss. Am. J. Reprod. Immunol. 2009, 61, 34–38. [Google Scholar] [CrossRef]
  198. Yenicesu, G.I.; Cetin, M.; Ozdemir, O.; Cetin, A.; Ozen, F.; Yenicesu, C.; Yildiz, C.; Kocak, N. A Prospective case–control study analyzes 12 thrombophilic gene mutations in Turkish couples with recurrent pregnancy loss. Am. J. Reprod. Immunol. 2010, 63, 126–136. [Google Scholar] [CrossRef] [PubMed]
  199. Ozornek, H.; Ergin, E.; Jeyendran, R.S.; Ozay, A.T.; Pillai, D.; Coulam, C. Is apolipoprotien E codon 112 polymorphisms associated with recurrent pregnancy loss? Am. J. Reprod. Immunol. 2010, 64, 87–92. [Google Scholar] [CrossRef]
  200. Ergin, E.; Jeyendran, R.S.; Özörnek, H.; Alev, Ö.; Pillai, M.D.; Coulam, C. Apolipoprotein E codon 112 polymorphisms is associated with recurrent pregnancy loss. Fertil. Steril. 2009, 92, S115. [Google Scholar] [CrossRef]
  201. Vasunilashorn, S.; Finch, C.E.; Crimmins, E.M.; Vikman, S.A.; Stieglitz, J.; Gurven, M.; Kaplan, H.; Allayee, H. Inflammatory gene variants in the Tsimane, an indigenous Bolivian population with a high infectious load. Biodemography Soc. Biol. 2011, 57, 33–52. [Google Scholar] [CrossRef]
  202. Trumble, B.C.; Stieglitz, J.; Blackwell, A.D.; Allayee, H.; Beheim, B.; Finch, C.E.; Gurven, M.; Kaplan, H. Apolipoprotein E4 is associated with improved cognitive function in Amazonian forager-horticulturalists with a high parasite burden. FASEB J. 2017, 31, 1508–1515. [Google Scholar] [CrossRef]
  203. Tzourio, C.; Arima, H.; Harrap, S.; Anderson, C.; Godin, O.; Woodward, M.; Neal, B.; Bousser, M.-G.; Chalmers, J.; Cambien, F.; et al. APOE genotype, ethnicity, and the risk of cerebral hemorrhage. Neurology 2008, 70, 1322–1328. [Google Scholar] [CrossRef]
  204. Zhang, M.; Gu, W.; Qiao, S.; Zhu, E.; Zhao, Q.; Lv, S. Apolipoprotein E gene polymorphism and risk for coronary heart disease in the Chinese population: A meta-analysis of 61 studies including 6634 cases and 6393 controls. PLoS ONE 2014, 9, e9546. [Google Scholar] [CrossRef]
  205. Tang, M.-X.; Stern, Y.; Marder, K.; Bell, K.; Gurland, B.; Lantigua, R.; Andrews, H.; Feng, L.; Tycko, B.; Mayeux, R. The APOE-∊4 allele and the risk of Alzheimer disease among African Americans, Whites, and Hispanics. JAMA 1998, 279, 751–755. [Google Scholar] [CrossRef]
  206. Wright, R.O.; Hu, H.; Silverman, E.K.; Tsaih, S.W.; Schwartz, J.; Bellinger, D.; Palazuelos, E.; Weiss, S.T.; Hernandez-Avila, M. Apolipoprotein E genotype predicts 24-Month Bayley scales infant development score. Pediatr. Res. 2003, 54, 819–825. [Google Scholar] [CrossRef]
  207. Becher, J.-C.; Bell, J.E.; McIntosh, N.; Keeling, J.W. Distribution of apolipoprotein E alleles in a Scottish healthy newborn population. Neonatology 2005, 88, 164–167. [Google Scholar] [CrossRef]
  208. Becher, J.; Keeling, J.W.; McIntosh, N.; Wyatt, B.; Bell, J. The distribution of apolipoprotein E alleles in Scottish perinatal deaths. J. Med. Genet. 2006, 43, 414–418. [Google Scholar] [CrossRef]
  209. Becher, J.-C.; Keeling, J.W.; Bell, J.; Wyatt, B.; McIntosh, N. Apolipoprotein E e4 and its prevalence in early childhood death due to sudden infant death syndrome or to recognised causes. Early Hum. Dev. 2008, 84, 549–554. [Google Scholar] [CrossRef]
  210. Farrer, L.A.; Cupples, L.A.; Haines, J.L.; Hyman, B.; Kukull, W.A.; Mayeux, R.; Myers, R.H.; Pericak-Vance, M.A.; Risch, N.; van Duijn, C.M. Effects of age, sex, and ethnicity on the association between apolipoprotein E genotype and Alzheimer disease: A meta-analysis. JAMA 1997, 278, 1349–1356. [Google Scholar] [CrossRef]
  211. Bennet, A.M.; Angelantonio, E.D.; Ye, Z.; Wensley, F.; Dahlin, A.; Ahlbom, A.; Keavney, B.; Collins, R.; Wiman, B.; de Faire, U.; et al. Association of apolipoprotein E genotypes with lipid levels and coronary risk. JAMA 2007, 298, 1300–1311. [Google Scholar] [CrossRef]
  212. Rougeron, V.; Woods, C.M.; Tiedje, K.E.; Bodeau-Livinec, F.; Migot-Nabias, F.; Deloron, P.; Luty, A.J.F.; Fowkes, F.J.I.; Day, K.P. Epistatic interactions between apolipoprotein E and Hemoglobin S genes in regulation of Malaria Parasitemia. PLoS ONE 2013, 8, e76924. [Google Scholar] [CrossRef] [PubMed]
  213. Kulminski, A.M.; Raghavachari, N.; Arbeev, K.G.; Culminskaya, I.; Arbeeva, L.; Wu, D.; Ukraintseva, S.V.; Christensen, K.; Yashin, A.I. Protective role of the apolipoprotein E2 allele in age-related disease traits and survival: Evidence from the long life family study. Biogerontology 2016, 17, 893–905. [Google Scholar] [CrossRef]
  214. Konishi, K.; Bhat, V.; Banner, H.; Poirier, J.; Joober, R.; Bohbot, V.D. APOE2 is associated with spatial navigational strategies and increased gray matter in the hippocampus. Front. Hum. Neurosci. 2016, 10. [Google Scholar] [CrossRef]
  215. Cashdan, E. Biogeography of human infectious diseases: A global historical analysis. PLoS ONE 2014, 9. [Google Scholar] [CrossRef]
  216. Guernier, V.; Hochberg, M.E.; Guégan, J.-F. Ecology drives the worldwide distribution of human diseases. PLoS Biol. 2004, 2, e141. [Google Scholar] [CrossRef]
  217. Barger, S.W.; Harmon, A.D. Microglial activation by Alzheimer amyloid precursor protein and modulation by apolipoprotein E. Nature 1997, 388, 878–881. [Google Scholar] [CrossRef] [PubMed]
  218. Bonacina, F.; Coe, D.; Wang, G.; Longhi, M.P.; Baragetti, A.; Moregola, A.; Garlaschelli, K.; Uboldi, P.; Pellegatta, F.; Grigore, L.; et al. Myeloid apolipoprotein E controls dendritic cell antigen presentation and T cell activation. Nat. Commun. 2018, 9, 3083. [Google Scholar] [CrossRef]
  219. Kelly, M.E.; Clay, M.A.; Mistry, M.J.; Hsieh-Li, H.-M.; Harmony, J.A.K. Apolipoprotein E inhibition of proliferation of mitogen-activated T lymphocytes: Production of interleukin 2 with reduced biological activity. Cell. Immunol. 1994, 159, 124–139. [Google Scholar] [CrossRef]
  220. Picardi, A.; Gentilucci, U.V.; Bambacioni, F.; Galati, G.; Spataro, S.; Mazzarelli, C.; D’Avola, D.; Fiori, E.; Riva, E. Lower schooling, higher hepatitis C virus prevalence in Italy: An association dependent on age. J. Clin. Virol. 2007, 40, 168–170. [Google Scholar] [CrossRef]
  221. Corbo, R.M.; Scacchi, R.; Mureddu, L.; Mulas, G.; Alfano, G. Apolipoprotein E polymorphism in Italy investigated in native plasma by a simple polyacrylamide gel isoelectric focusing technique. Comparison with frequency data of other European populations. Ann. Hum. Genet. 1995, 59, 197–209. [Google Scholar] [CrossRef]
  222. Kuhlmann, I.; Minihane, A.M.; Huebbe, P.; Nebel, A.; Rimbach, G. Apolipoprotein E genotype and hepatitis C, HIV and herpes simplex disease risk: A literature review. Lipids Health Dis. 2010, 9, 8. [Google Scholar] [CrossRef]
  223. Soares, H.D.; Potter, W.Z.; Pickering, E.; Kuhn, M.; Immermann, F.W.; Shera, D.M.; Ferm, M.; Dean, R.A.; Simon, A.J.; Swenson, F.; et al. Biomarkers associated with the apolipoprotein E genotype and Alzheimer disease. Arch. Neurol. 2012, 69, 1310–1317. [Google Scholar] [CrossRef]
  224. Gale, S.C.; Gao, L.; Mikacenic, C.; Coyle, S.M.; Rafaels, N.; Murray, T.; Madenspacher, J.H.; Draper, D.W.; Ge, W.; Aloor, J.J.; et al. APOε 4 is associated with enhanced in vivo innate immune responses in humans. J. Allergy Clin. Immunol. 2014, 134, 127–134. [Google Scholar] [CrossRef]
  225. Oosten, M.V.; Rensen, P.C.N.; Amersfoort, E.S.V.; Eck, M.V.; Dam, A.-M.V.; Brevé, J.J.P.; Vogel, T.; Panet, A.; Berkel, T.J.C.V.; Kuiper, J. Apolipoprotein E protects against bacterial lipopolysaccharide-induced lethality a new therapeutic approach to treat gram-negative sepsis. J. Biol. Chem. 2001, 276, 8820–8824. [Google Scholar] [CrossRef]
  226. Liu, B.; Zhang, Y.; Wang, R.; An, Y.; Gao, W.; Bai, L.; Li, Y.; Zhao, S.; Fan, J.; Liu, E. Western diet feeding influences gut microbiota profiles in apoE knockout mice. Lipids Health Dis. 2018, 17, 159. [Google Scholar] [CrossRef]
  227. Kasahara, K.; Tanoue, T.; Yamashita, T.; Yodoi, K.; Matsumoto, T.; Emoto, T.; Mizoguchi, T.; Hayashi, T.; Kitano, N.; Sasaki, N.; et al. Commensal bacteria at the crossroad between cholesterol homeostasis and chronic inflammation in atherosclerosis. J. Lipid Res. 2017, 58, 519–528. [Google Scholar] [CrossRef]
  228. Saita, D.; Ferrarese, R.; Foglieni, C.; Esposito, A.; Canu, T.; Perani, L.; Ceresola, E.R.; Visconti, L.; Burioni, R.; Clementi, M.; et al. Adaptive immunity against gut microbiota enhances apoE-mediated immune regulation and reduces atherosclerosis and western-diet-related inflammation. Sci. Rep. 2016, 6, 29353. [Google Scholar] [CrossRef]
  229. Schneeberger, M.; Everard, A.; Gómez-Valadés, A.G.; Matamoros, S.; Ramírez, S.; Delzenne, N.M.; Gomis, R.; Claret, M.; Cani, P.D. Akkermansia muciniphila inversely correlates with the onset of inflammation, altered adipose tissue metabolism and metabolic disorders during obesity in mice. Sci. Rep. 2015, 5, 16643. [Google Scholar] [CrossRef]
  230. Li, J.; Lin, S.; Vanhoutte, P.M.; Woo, C.W.; Xu, A. Akkermansia Muciniphila protects against atherosclerosis by preventing metabolic endotoxemia-induced inflammation in ApoE−/− mice. Circulation 2016, 133, 2434–2446. [Google Scholar] [CrossRef]
  231. Zhao, S.; Liu, W.; Wang, J.; Shi, J.; Sun, Y.; Wang, W.; Ning, G.; Liu, R.; Hong, J. Akkermansia muciniphila improves metabolic profiles by reducing inflammation in chow diet-fed mice. J. Mol. Endocrinol. 2017, 58, 1–14. [Google Scholar] [CrossRef] [PubMed]
  232. Yashin, A.I.; Wu, D.; Arbeev, K.G.; Ukraintseva, S.V. Polygenic effects of common single-nucleotide polymorphisms on life span: When association meets causality. Rejuvenation Res. 2012, 15, 381–394. [Google Scholar] [CrossRef] [PubMed]
  233. Boyle, E.A.; Li, Y.I.; Pritchard, J.K. An expanded view of complex traits: From polygenic to omnigenic. Cell 2017, 169, 1177–1186. [Google Scholar] [CrossRef] [PubMed]
  234. Gnecchi-Ruscone, G.A.; Abondio, P.; De Fanti, S.; Sarno, S.; Sherpa, M.G.; Sherpa, P.T.; Marinelli, G.; Natali, L.; Di Marcello, M.; Peluzzi, D.; et al. Evidence of polygenic adaptation to high altitude from Tibetan and Sherpa genomes. Genome Biol. Evol. 2018, 10, 2919–2930. [Google Scholar] [CrossRef]
  235. McClellan, J.; King, M.-C. Genetic heterogeneity in human disease. Cell 2010, 141, 210–217. [Google Scholar] [CrossRef] [PubMed]
  236. Wang, B.; Bao, S.; Zhang, Z.; Zhou, X.; Wang, J.; Fan, Y.; Zhang, Y.; Li, Y.; Chen, L.; Jia, Y.; et al. A rare variant in MLKL confers susceptibility to ApoE ɛ4-negative Alzheimer’s disease in Hong Kong Chinese population. Neurobiol. Aging 2018, 68, 160.e1–160.e7. [Google Scholar] [CrossRef] [PubMed]
  237. Evolutionary Thinking in Medicine: From Research to Policy and Practice; Alvergne, A.; Jenkinson, C.; Faurie, C. (Eds.) Springer International Publishing: Basel, Switzerland, 2016. [Google Scholar]
  238. Allentoft, M.E.; Sikora, M.; Sjögren, K.-G.; Rasmussen, S.; Rasmussen, M.; Stenderup, J.; Damgaard, P.B.; Schroeder, H.; Ahlström, T.; Vinner, L.; et al. Population genomics of Bronze Age Eurasia. Nature 2015, 522, 167–172. [Google Scholar] [CrossRef]
  239. Olalde, I.; Brace, S.; Allentoft, M.E.; Armit, I.; Kristiansen, K.; Booth, T.; Rohland, N.; Mallick, S.; Szécsényi-Nagy, A.; Mittnik, A.; et al. The Beaker phenomenon and the genomic transformation of northwest Europe. Nature 2018, 555, 190–196. [Google Scholar] [CrossRef] [PubMed]
  240. Olalde, I.; Schroeder, H.; Sandoval-Velasco, M.; Vinner, L.; Lobón, I.; Ramirez, O.; Civit, S.; García Borja, P.; Salazar-García, D.C.; Talamo, S.; et al. A common genetic origin for early farmers from Mediterranean Cardial and Central European LBK cultures. Mol. Biol. Evol. 2015, 32, 3132–3142. [Google Scholar] [CrossRef]
  241. Olalde, I.; Allentoft, M.E.; Sánchez-Quinto, F.; Santpere, G.; Chiang, C.W.K.; DeGiorgio, M.; Prado-Martinez, J.; Rodríguez, J.A.; Rasmussen, S.; Quilez, J.; et al. Derived immune and ancestral pigmentation alleles in a 7000-year-old Mesolithic European. Nature 2014, 507, 225–228. [Google Scholar] [CrossRef]
  242. Mathieson, I.; Alpaslan-Roodenberg, S.; Posth, C.; Szécsényi-Nagy, A.; Rohland, N.; Mallick, S.; Olalde, I.; Broomandkhoshbacht, N.; Candilio, F.; Cheronet, O.; et al. The genomic history of southeastern Europe. Nature 2018, 555, 197–203. [Google Scholar] [CrossRef] [PubMed]
  243. Mathieson, I.; Lazaridis, I.; Rohland, N.; Mallick, S.; Patterson, N.; Roodenberg, S.A.; Harney, E.; Stewardson, K.; Fernandes, D.; Novak, M.; et al. Genome-wide patterns of selection in 230 ancient Eurasians. Nature 2015, 528, 499–503. [Google Scholar] [CrossRef] [PubMed]
  244. Jones, E.R.; Gonzalez-Fortes, G.; Connell, S.; Siska, V.; Eriksson, A.; Martiniano, R.; McLaughlin, R.L.; Gallego Llorente, M.; Cassidy, L.M.; Gamba, C.; et al. Upper Palaeolithic genomes reveal deep roots of modern Eurasians. Nat. Commun. 2015, 6, 8912. [Google Scholar] [CrossRef]
  245. Haak, W.; Lazaridis, I.; Patterson, N.; Rohland, N.; Mallick, S.; Llamas, B.; Brandt, G.; Nordenfelt, S.; Harney, E.; Stewardson, K.; et al. Massive migration from the steppe was a source for Indo-European languages in Europe. Nature 2015, 522, 207–211. [Google Scholar] [CrossRef] [PubMed]
  246. Lipson, M.; Szécsényi-Nagy, A.; Mallick, S.; Pósa, A.; Stégmár, B.; Keerl, V.; Rohland, N.; Stewardson, K.; Ferry, M.; Michel, M.; et al. Parallel palaeogenomic transects reveal complex genetic history of early European farmers. Nature 2017, 551, 368–372. [Google Scholar] [CrossRef] [PubMed]
  247. Lazaridis, I.; Patterson, N.; Mittnik, A.; Renaud, G.; Mallick, S.; Kirsanow, K.; Sudmant, P.H.; Schraiber, J.G.; Castellano, S.; Lipson, M.; et al. Ancient human genomes suggest three ancestral populations for present-day Europeans. Nature 2014, 513, 409–413. [Google Scholar] [CrossRef] [PubMed]
  248. Lazaridis, I.; Mittnik, A.; Patterson, N.; Mallick, S.; Rohland, N.; Pfrengle, S.; Furtwängler, A.; Peltzer, A.; Posth, C.; Vasilakis, A.; et al. Genetic origins of the Minoans and Mycenaeans. Nature 2017, 548, 214–218. [Google Scholar] [CrossRef] [PubMed]
  249. Schiffels, S.; Haak, W.; Paajanen, P.; Llamas, B.; Popescu, E.; Loe, L.; Clarke, R.; Lyons, A.; Mortimer, R.; Sayer, D.; et al. Iron Age and Anglo-Saxon genomes from East England reveal British migration history. Nat. Commun. 2016, 7, 10408. [Google Scholar] [CrossRef]
  250. Martiniano, R.; Caffell, A.; Holst, M.; Hunter-Mann, K.; Montgomery, J.; Müldner, G.; McLaughlin, R.L.; Teasdale, M.D.; van Rheenen, W.; Veldink, J.H.; et al. Genomic signals of migration and continuity in Britain before the Anglo-Saxons. Nat. Commun. 2016, 7, 10326. [Google Scholar] [CrossRef]
  251. Broushaki, F.; Thomas, M.G.; Link, V.; López, S.; van Dorp, L.; Kirsanow, K.; Hofmanová, Z.; Diekmann, Y.; Cassidy, L.M.; Díez-del-Molino, D.; et al. Early Neolithic genomes from the eastern Fertile Crescent. Science 2016, 353, 499–503. [Google Scholar] [CrossRef] [PubMed]
  252. Cassidy, L.M.; Martiniano, R.; Murphy, E.M.; Teasdale, M.D.; Mallory, J.; Hartwell, B.; Bradley, D.G. Neolithic and Bronze Age migration to Ireland and establishment of the insular Atlantic genome. Proc. Natl. Acad. Sci. USA 2016, 113, 368–373. [Google Scholar] [CrossRef] [PubMed]
  253. Raghavan, M.; Skoglund, P.; Graf, K.E.; Metspalu, M.; Albrechtsen, A.; Moltke, I.; Rasmussen, S.; Stafford, T.W., Jr.; Orlando, L.; Metspalu, E.; et al. Upper Palaeolithic Siberian genome reveals dual ancestry of Native Americans. Nature 2014, 505, 87–91. [Google Scholar] [CrossRef] [PubMed]
  254. Fu, Q.; Li, H.; Moorjani, P.; Jay, F.; Slepchenko, S.M.; Bondarev, A.A.; Johnson, P.L.F.; Aximu-Petri, A.; Prüfer, K.; de Filippo, C.; et al. Genome sequence of a 45,000-year-old modern human from western Siberia. Nature 2014, 514, 445–449. [Google Scholar] [CrossRef] [PubMed]
  255. Günther, T.; Valdiosera, C.; Malmström, H.; Ureña, I.; Rodriguez-Varela, R.; Sverrisdóttir, Ó.O.; Daskalaki, E.A.; Skoglund, P.; Naidoo, T.; Svensson, E.M.; et al. Ancient genomes link early farmers from Atapuerca in Spain to modern-day Basques. Proc. Natl. Acad. Sci. USA 2015, 112, 11917–11922. [Google Scholar] [CrossRef] [PubMed]
  256. Hofmanová, Z.; Kreutzer, S.; Hellenthal, G.; Sell, C.; Diekmann, Y.; Díez-del-Molino, D.; van Dorp, L.; López, S.; Kousathanas, A.; Link, V.; et al. Early farmers from across Europe directly descended from Neolithic Aegeans. Proc. Natl. Acad. Sci. USA 2016, 113, 6886–6891. [Google Scholar] [CrossRef] [PubMed]
  257. Kılınç, G.M.; Omrak, A.; Özer, F.; Günther, T.; Büyükkarakaya, A.M.; Bıçakçı, E.; Baird, D.; Dönertaş, H.M.; Ghalichi, A.; Yaka, R.; et al. The Demographic development of the First Farmers in Anatolia. Curr. Biol. 2016, 26, 2659–2666. [Google Scholar] [CrossRef] [PubMed]
  258. Keller, A.; Graefen, A.; Ball, M.; Matzas, M.; Boisguerin, V.; Maixner, F.; Leidinger, P.; Backes, C.; Khairat, R.; Forster, M.; et al. New insights into the Tyrolean Iceman’s origin and phenotype as inferred by whole-genome sequencing. Nat. Commun. 2012, 3, 698. [Google Scholar] [CrossRef] [PubMed]
  259. Nesse, R.M. Evolution: Medicine’s most basic science. Lancet 2008, 372, S21–S27. [Google Scholar] [CrossRef]
  260. Nesse, R.M. Ten questions for evolutionary studies of disease vulnerability. Evol. Appl. 2011, 4, 264–277. [Google Scholar] [CrossRef] [PubMed]
  261. Nesse, R.M.; Bergstrom, C.T.; Ellison, P.T.; Flier, J.S.; Gluckman, P.; Govindaraju, D.R.; Niethammer, D.; Omenn, G.S.; Perlman, R.L.; Schwartz, M.D.; et al. Making evolutionary biology a basic science for medicine. Proc. Natl. Acad. Sci. USA 2010, 107, 1800–1807. [Google Scholar] [CrossRef] [PubMed]
  262. Wells, J.C.K.; Nesse, R.M.; Sear, R.; Johnstone, R.A.; Stearns, S.C. Evolutionary public health: Introducing the concept. Lancet 2017, 390, 500–509. [Google Scholar] [CrossRef]
  263. Randolph, M.; Nesse, M.D.; George, C.W. Why We Get Sick. Available online: https://www.penguinrandomhouse.com/books/120768/why-we-get-sick-by-randolph-m-nesse/9780679746744 (accessed on 1 March 2019).
Figure 1. Polymorphisms underlying the three main APOE variants in humans. (A) Chromosome location, gene structure, identity of the mutating sites in the gene, and the corresponding mutating residues in the context of the protein structure. In yellow, it is indicated as the receptor-binding region in helix 4 and, in green, it is the lipid-binding region in the C-terminal domain. Red and black dots indicate the genetic variants in APOE and their position in the genomic and protein sequences, respectively. (B) Table reporting the haplotypes and corresponding residue combination associated to each APOE allele.
Figure 1. Polymorphisms underlying the three main APOE variants in humans. (A) Chromosome location, gene structure, identity of the mutating sites in the gene, and the corresponding mutating residues in the context of the protein structure. In yellow, it is indicated as the receptor-binding region in helix 4 and, in green, it is the lipid-binding region in the C-terminal domain. Red and black dots indicate the genetic variants in APOE and their position in the genomic and protein sequences, respectively. (B) Table reporting the haplotypes and corresponding residue combination associated to each APOE allele.
Genes 10 00222 g001
Figure 2. Frequency distribution of APOE alleles in 82 countries. Data from the 1000 Genome Project have been integrated with those published in Singh et al. 2006. (A) Frequency distribution of the ε2 variant. (B) Frequency distribution of the ε3 variant. (C) Frequency distribution of the ε4 variant.
Figure 2. Frequency distribution of APOE alleles in 82 countries. Data from the 1000 Genome Project have been integrated with those published in Singh et al. 2006. (A) Frequency distribution of the ε2 variant. (B) Frequency distribution of the ε3 variant. (C) Frequency distribution of the ε4 variant.
Genes 10 00222 g002
Figure 3. Distribution and approximate age of the analyzed ancient samples. Those coming from the same location and belonging to the same culture have been clustered together and share the same label. The number of grouped individuals is given in brackets.
Figure 3. Distribution and approximate age of the analyzed ancient samples. Those coming from the same location and belonging to the same culture have been clustered together and share the same label. The number of grouped individuals is given in brackets.
Genes 10 00222 g003

Share and Cite

MDPI and ACS Style

Abondio, P.; Sazzini, M.; Garagnani, P.; Boattini, A.; Monti, D.; Franceschi, C.; Luiselli, D.; Giuliani, C. The Genetic Variability of APOE in Different Human Populations and Its Implications for Longevity. Genes 2019, 10, 222. https://doi.org/10.3390/genes10030222

AMA Style

Abondio P, Sazzini M, Garagnani P, Boattini A, Monti D, Franceschi C, Luiselli D, Giuliani C. The Genetic Variability of APOE in Different Human Populations and Its Implications for Longevity. Genes. 2019; 10(3):222. https://doi.org/10.3390/genes10030222

Chicago/Turabian Style

Abondio, Paolo, Marco Sazzini, Paolo Garagnani, Alessio Boattini, Daniela Monti, Claudio Franceschi, Donata Luiselli, and Cristina Giuliani. 2019. "The Genetic Variability of APOE in Different Human Populations and Its Implications for Longevity" Genes 10, no. 3: 222. https://doi.org/10.3390/genes10030222

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

Article Metrics

Back to TopTop