Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Aug 24.
Published in final edited form as: Neurosci Lett. 2019 Jun 7;708:134306. doi: 10.1016/j.neulet.2019.134306

Peripheral Versus Central Nervous System APOE in Alzheimer’s Disease: Interplay across the Blood-Brain Barrier

Dustin Chernick 1, Stephanie Ortiz-Valle 1, Angela Jeong 2, Wenhui Qu 3, Ling Li 1,2,3,*
PMCID: PMC6693948  NIHMSID: NIHMS1532351  PMID: 31181302

Abstract

The apolipoprotein E (APOE) ε4 allele has been demonstrated as the preeminent genetic risk factor for late onset Alzheimer’s disease (AD), which comprises greater than 90% of all AD cases. The discovery of the connection between different APOE genotypes and AD risk in the early 1990s spurred three decades of intense and comprehensive research into the function of APOE in the normal and diseased brain. The importance of APOE in the periphery has been well established, due to its pivotal role in maintaining cholesterol homeostasis and cardiovascular health. The influence of vascular factors on brain function and AD risk has been extensively studied in recent years. As a major apolipoprotein regulating multiple molecular pathways beyond its canonical lipid-related functions in the periphery and the central nervous system, APOE represents a critical link between the two compartments, and may influence AD risk from both perspectives. This review discusses recent advances in understanding the different functions of APOE in the periphery and in the brain, as well as highlighting several promising APOE-targeted therapeutic strategies for AD.

Keywords: Apolipoprotein E, Periphery, Brain, Alzheimer’s disease pathogenesis, Inflammation, Blood-brain barrier, Cognition, APOE-targeted therapeutic strategies

Introduction

Apolipoprotein E (APOE) is a 34 kDa glycoprotein that is synthesized primarily in the liver and in the brain. ApoE is incorporated into lipoprotein particles in the plasma in the periphery, as well as in the cerebrospinal fluid (CSF) and in the interstitial fluid (ISF) of brain parenchyma in the central nervous system (CNS) [1, 2]. The canonical function of apoE is in transporting cholesterol and other lipids, mediated by cell surface apoE receptors, although APOE also plays important roles in immunomodulation, synaptic plasticity, signal transduction, and proteostasis [3, 4].

The APOE gene is polymorphic in humans, consisting of three alleles (ε2, ε3, and ε4), which leads to the production of three distinct isoforms of APOE protein: APOE2, APOE3, and APOE4. The three APOE isoforms differ by one single amino acid residue at position 112 or 158: APOE2 has cysteine residues at both positions, APOE3 has a cysteine residue at 112 and an arginine residue at 158, and APOE4 has arginine residues at both positions [5, 6]. These differences alter the structure and physiological function of APOE. APOE2 may elicit type III hyperlipoproteinemia and APOE4 increases the risk for hypercholesterolemia and atherosclerosis [7].

Interest in APOE in the CNS escalated greatly when it was discovered that the APOE4 allele is associated with a significant increase in the risk for late-onset Alzheimer’s disease (AD), while the presence of the APOE2 allele protects against the development of AD [810]. The allele frequency of APOE4 is approximately 15% in the general population; however, it is enriched to ~ 40% in AD patients. APOE3, the most common allele, has a frequency of approximately 77%, and is considered to be neutral in regards to neurodegeneration, while the APOE2 allele is less common in the general population (~ 8% allele frequency) [810].

Circulating lipoproteins are unable to cross the blood brain barrier; however, small high-density lipoprotein (HDL) particles can enter the brain, stimulating interest in identifying the relevance of plasma HDL and associated apolipoproteins in brain function [11, 12]. This review aims to capture recent findings on the role of both peripheral and brain APOE in the pathogenesis of AD and highlight emerging APOE-targeted therapeutics for AD.

Two pools of APOE

The most compelling support for two distinct pools of APOE in the human body came first from a liver transplantation study in 1991 [13]. It was found that in patients after liver transplantation, the APOE isoform present in the plasma almost completely converted to that of the donor whereas the APOE isoform in the brain did not change. This study demonstrated that APOE in the periphery is primarily produced by the liver and that the protein does not cross the blood brain barrier (BBB) [13]. In the brain, APOE is constitutively produced by glial cells (astrocytes, oligodendrocytes, and microglia), pericytes, vascular smooth muscle cells, choroid plexus, and limitedly by neurons under stress conditions [1416]. In the periphery, although the major source of APOE is from hepatocytes in the liver, APOE is also expressed by monocytes and macrophages [1, 17, 18]. Concentrations of APOE differ highly between the two compartments; APOE levels are around 50–70 μg/mL in the plasma, whereas they are in the range of 3–10 μg/mL in the CSF [19, 20].

The regulation of APOE expression is tissue- and cell-specific and is controlled by several transcription factors and regulatory elements, as recently reviewed [18]. The expression of APOE can be induced by environmental factors including cytokines, hormones, or lipids. In the periphery, APOE expression in the hepatocytes and macrophages requires the binding of specific distal enhancers, as well as common proximal enhancer elements, to the promoter region. Notably, although the amount of APOE expression is low in macrophages, local production of APOE by macrophage exerts dramatic protection against hypercholesterolemia-induced atherosclerosis [21, 22]. In the CNS, glial APOE expression is regulated by specific transcriptional factors through distinct binding sites on its promoter [18]. Neuronal APOE expression is uniquely regulated, in which an intron-containing APOE splice variant is produced but not translated; upon injury, this splice variant is further processed to produce normal mRNA for translation to protein [23]. In addition, recent studies have indicated the expression of APOE is also subjected to certain epigenetic controls [18].

The posttranslational modification of APOE in the periphery and CNS is also differentially regulated. Newly produced APOE follows the classical secretory pathway. Glycosylation and sialyation of APOE occurs in the Golgi and are uniformly present in APOE secreted from hepatocytes and macrophages [24]. However, APOE in plasma typically lacks carbohydrate molecules [18, 25]. In contrast, APOE in CSF remains heavily glycosylated and sialylated, consistent with the form of APOE secreted from astrocytes. The sialic acid groups reduce the isoelectric point of apoE, and glycosylated apoE is less soluble in Tris buffered saline (TBS) containing the detergent Triton-X100 than in TBS alone, suggesting that this post-translational modification may influence the ability of APOE to bind lipids [25]. APOE4 was more commonly sialylated, and more commonly found in the TBS-soluble fraction, and less commonly in the fraction containing detergent, than APOE3, suggesting that this post-translational modification may underlie the reduced lipid binding observed for APOE4.

Recent studies of APOE levels in CSF, brain parenchyma, and plasma of individuals with different APOE genotypes have provided further evidence of distinct peripheral and central APOE pools in the context of AD. It has been shown that the correlation between CSF and plasma APOE levels is very low [26]. Importantly, APOE4 carriers have a decreased level of APOE, whereas APOE2 carriers have an increased level of APOE in CSF and plasma compared with APOE3 carriers [26]. This study further showed that CSF but not plasma APOE levels are associated with CSF Aβ42 levels. However, another study found no differences in CSF APOE levels between APOE genotypes but showed that APOE4 carriers exhibit significantly decreased plasma APOE levels, resulting from a specific decrease in APOE4 concentrations [27], and suggested that the life-long decrease in plasma APOE may predispose APOE4 carriers to AD. Another recent study reported a robust imbalance in APOE isoform distribution in APOE3/APOE4 heterozygous individuals [28]. In CSF and cortical tissue, APOE4 levels were higher than APOE3 levels, whereas in plasma the relationship was reversed; CSF and plasma APOE4/APOE3 isoform ratios were negatively correlated between paired plasma and CSF samples. These observations indicate that APOE in the periphery and CNS exhibits different isoform dependent turnover rates. Plasma APOE turnover is isoform-dependent with a rank order of APOE4 > APOE3 > APOE2. In contrast, CNS APOE turnover is three- to six- fold slower and is isoform independent. The study also showed that the level of APOE in CSF increases with age and that amyloid deposition is associated with further elevation of APOE3, but a decrease of APOE4 in the CSF [28]. Further, APOE levels in the CSF were associated with CSF amyloid levels, whereas plasma APOE levels lacked this correlation. However, other recent large studies have shown that low plasma/serum APOE levels are associated with an increased risk of AD/dementia, independent of APOE genotype, in general population [2931].

Taken together, these studies demonstrate that peripheral and CNS APOE are separate pools and are distinctly regulated. Emerging evidence indicates peripheral APOE, as well as CNS APOE, influences the progression of AD directly or indirectly through multiple pathways as shown in Fig. 1 and discussed below.

Fig. 1.

Fig. 1.

A schematic to illustrate that APOE in the periphery and in the brain regulates multiple molecular pathways pertinent to the pathogenesis of AD. See the main text for details. Aβ: amyloid-β; ABCA1: ATP-binding cassette transporter A1; APOE or E: apolipoprotein E; BBB, blood-brain barrier; HDL, high-density lipoprotein; VLDL: very low-density lipoprotein; LDLR, low-density lipoprotein receptor, LRP1, low-density lipoprotein receptor-related protein 1; TREM2: triggering receptor expressed on myeloid cells 2.

APOE in cholesterol/lipid metabolism

APOE binds primarily to the low-density lipoprotein (LDL) receptor family, which plays crucial roles in cholesterol homeostasis. APOE3 and APOE2 are more selective to HDL, while APOE4 preferentially binds to triglyceride-rich, very low-density lipoproteins (VLDL), leading to downregulation of LDL receptors and thus reduced clearance of LDL and increased plasma cholesterol levels [7]. These differences in affinity lead to increased risk of atherosclerosis and stroke among APOE4 carriers, as well as increased risk of type III hyperlipoproteinemia in those carrying APOE2, especially in homozygous APOE2 individuals because of defective binding of APOE2 to LDL receptors [7]. Some studies indicate that APOE2-induced type III hyperlipoproteinemia only occurs in the presence of other conditions such as diabetes, hypothyroidism, and obesity, which leads to fewer LDL receptors, limiting the ability of APOE2 to modulate cholesterol clearance [7]. The critical importance of APOE in cholesterol/lipid metabolism in the periphery was demonstrated by the phenotypes of ApoE-knockout mice, which include the spontaneous development of severe hypercholesterolemia and atherosclerosis [32, 33].

Cholesterol homeostasis is integral to normal brain function. While the brain comprises approximately 2% of the total body weight, as much as 25% of all cholesterol in the body is found in the brain [34]. APOE, as a major apolipoprotein produced primarily by astrocytes, plays a pivotal role in maintaining the cholesterol/lipid homeostasis in the brain [2, 35]. Once secreted, APOE binds cholesterol and phospholipids through interactions with the ATP-binding cassette transporter A1 (ABCA1), forming HDL-like particles in ISF and CSF. APOE mediates the interactions of APOE-containing lipoproteins and lipid complexes with cell surface heparin sulfate proteoglycan (HSPG) and cell membrane associated receptors, including the LDL receptor and the LDL receptor-related protein (LRP), promoting cellular uptake and redistribution of cholesterol/lipids, storage, or conversion of cholesterol to oxysterols for clearance through the blood-brain barrier [2, 35]. Importantly, APOE isoforms exhibit differential abilities of binding/transporting cholesterol and phospholipids (APOE2>APOE3>APOE4) [36, 37]. Consequently, APOE4 is poorly lipidated compared with APOE2 and APOE3 [16, 3841]. Intriguingly, a recent study shows APOE isoform-dependent effects on homeostasis of specific phospholipids, such as phosphoinositol biphosphate (PIP2) [42]. Such differences in lipidation and cholesterol/lipid homeostasis-maintaining capacity of APOE isoforms may account for APOE isoform-specific association with AD [4, 43, 44].

APOE in β-amyloidosis and tau pathology

The role of APOE in β-amyloidosis has been extensively studied and reviewed, in particular on Aβ aggregation and clearance [3, 16, 45]. Recent evidence highlights the role of APOE4 in driving early pathological accumulation/seeding of Aβ in the brain of AD model mice [46, 47]. In addition, a recent study shows experimental evidence that cholesterol plays a catalytic role in Aβ aggregation and that lipid membranes containing cholesterol increases the rate of Aβ aggregation [48]. Thus, the influence on cholesterol trafficking is a possible driving mechanism for the role of APOE in early pathogenesis. Notably, cholesterol-independent effects of APOE on APP/Aβ metabolism were shown recently in embryonic stem cell-derived human neurons [49]. In this study, APOE2, APOE3, and APOE4, either with or without cholesterol, exhibited differential stimulation on APP expression and Aβ production, with a potency rank order of APOE4 > APOE3 > APOE2. Further, it was revealed that APOE exerted those effects through a signal transduction pathway involving the activation of a non-canonical MAP kinase cascade [49].

The gain of toxic effects for APOE4 has been demonstrated recently using induced pluripotent stem cell (iPSC)-derived human neurons. APOE4-expressing neurons exhibited elevated Aβ production, tau phosphorylation, and GABAergic neuron atrophy, when compared to those isogenic neurons expressing APOE3 [50]. Highlighting the unique influence of APOE on different cell types, another similar study evaluated the role of APOE4 in neurons and glial cells. The authors found that transcriptional alterations associated with APOE4 were most commonly related to synaptic function in neurons, lipid metabolism in astrocytes, and immune activity in microglia [51]. While APOE4 neurons released greater levels of Aβ into the culture medium, microglia and astrocytes failed to clear the toxic peptide as effectively as their isogenic APOE3-expressing neurons. The use of iPSCs from human patients is a significant advance in the field of AD research, which offers unique insights into the role of APOE isoforms in human pathology. The use of iPSCs, differentiated to take on peripheral cell phenotypes, in particular vascular cells, may improve our understanding of the role of peripheral APOE in the pathological progression of AD as well.

Furthermore, the Aβ-independent role of APOE in tau pathology and tau-mediated neurodegeneration was recently demonstrated in animal models [52]. Compared with APOE2 and APOE3, APOE4 caused significant elevation of tau levels in the brain of human tau-expressing mice, followed by a marked increase of brain atrophy and neuroinflammation, whereas ApoE-deficient mice were protected from these changes. Further, the presence of the APOE4 allele was associated with more severe regional neurodegeneration in individuals with a sporadic primary tauopathy [52]. Consistent with the role of APOE4 in promoting tau pathology, APOE4-associated phospholipid dysregulation was shown to contribute to the development of tau hyperphosphorylation after traumatic brain injury [53]. In addition, APOE may also affect tau pathology through its effects in the periphery. It has been shown that in young healthy patients carrying the APOE4 allele, circulating peripheral lymphocytes express elevated levels of glycogen synthase kinase 3β (GSK3β) and phosphorylated tau, which were associated with subjective cognitive impairment [54]. In contrast, a recent study reports the association of APOE2, not APOE4, with increased tau pathology in primary tauopathy [55]. It is worth noting that the association of APOE4 or APOE2 with tau pathology may depend on the presence of amyloid pathology [56]. Clearly, more studies are needed to clarify the role of each APOE isoform in regulating tau aggregation under physiological and pathological conditions. Whether peripheral APOE pools participate in these processes also requires further investigation.

APOE in systemic inflammation and neuroinflammation

Compelling evidence indicates that inflammation plays a crucial role in the pathogenesis of AD [57] and that APOE modulates immune/inflammatory functions in the periphery and in the CNS [5760]. Recent studies highlight the impact of both systemic inflammation and neuroinflammation, and their interactions with different APOE isoforms to the progression of AD.

C-reactive protein, a key player of systemic immune response, has been used to investigate the relationship between systemic inflammation and AD. While multiple AD-related genes are associated with the level of CRP [61], the relationship between serum CRP levels and risk of AD has been inconclusive [6267]. One major limitation in those earlier studies was the measurement of CRP only at one time point. Recently, in a large, longitudinal study of the Framingham Heart Study offspring cohort, multiple measurements of CRP were taken to define a chronic condition of low-grade inflammation at baseline, its interaction with APOE genotype, and its association with the incidence of AD [68]. The results demonstrate that chronic low-grade inflammation interacts with APOE4 to increase the risk of AD and accelerate the onset of AD in a pattern dependent on CRP levels. Importantly, the association between CRP levels and risk of AD for APOE4 carriers became even more significant in the absence of cardiovascular diseases. In addition, while the elevated CRP level was linked to high rates of mortality across all APOE genotypes, an increased risk of AD was observed only in APOE4 carriers. In agreement, another study of a large, ethnically diverse longitudinal cohort showed that among > 100 serum proteins analyzed, only CRP was found to be significantly associated with APOE4 to mediate its contribution to AD risk [69]. In animal models expressing human APOE isoforms, it has also been demonstrated that APOE4, Aβ, and peripheral inflammation interact to induce cerebrovascular damage and cognitive deficits [70]. Furthermore, microRNAs may be involved in mediating the impact of APOE on inflammation. It has been reported that the level of miR146a, a major regulator of innate immune function, is significantly reduced in both the brain and plasma of APOE4 targeted replacement mice, leading to heightened pro-inflammatory signaling centrally and systemically in the absence or presence of Aβ [71]. Together, these findings indicate that APOE4 interacts with chronic systemic inflammation to increase the risk of AD.

In the meantime, much progress has been made lately in understanding the contribution of innate immunity in the brain to the pathogenesis of AD, in particular since the discovery that mutations in the triggering receptor expressed on myeloid cells 2 (TREM2) were associated with increased risk of AD [72, 73]. The role of TREM2 in AD has been covered recently by excellent reviews [7477]. The focus here is on the APOE-TREM2 interactions and their roles in AD. Following the discovery of association between TREM2 mutations and AD, several studies have identified APOE as a ligand for TREM2 and the APOE-TREM2 interaction modulates phagocytic function of microglia [7882]. In particular, it has been shown that the APOE-TREM2 pathway drives the dysfunctional microglial phenotypes in neurodegenerative diseases including AD [80, 83]. Using RiboTag translational profiling approach, a recent study discovered that aging, amyloid, and tau pathology converge on APOE pathway in microglia [84]. In addition, plaque-associated APOE origins from microglia and requires TREM2, suggesting microglial APOE contributes to AD pathogenesis [85]. Consistently, deletion of ApoE reduces microglial activation and neuroinflammation in the presence of Aβ or tau pathology [52, 86]. Importantly, the immunomodulatory role of APOE in AD and neurodegeneration may differ depending on the stage of the disease [87]. Regarding different APOE isoforms, a small clinical study showed that the presence of APOE4 allele was necessary to develop AD in TREM2-R47H carriers. However, another study with a larger cohort showed that many TREM2-R47H carriers did develop pathologically confirmed AD in the absence of APOE4 allele. Although an in vitro study found no evidence for APOE isoform-dependent differences in binding to or signaling through TREM2 [82], whether APOE isoforms differentially interact with TREM2 in vivo awaits further investigation. In addition, since astrocytes are the major source of APOE in the brain, it is unclear whether there are any structural and functional differences in APOE from astrocytes and microglia.

While the majority of research on TREM2 has focused on microglia in the brain, it is noteworthy that TREM2 is also expressed on myeloid cells in the periphery, including dendritic cells, mononuclear cells, and tissue macrophages [74, 88]. Recent studies have reported that TREM2 expression is increased in the peripheral blood mononuclear cells in AD patients [89], and in MCI patients who later progress to AD [90]. Whether peripheral pools of APOE and TREM2 interact in the plasma or at the BBB remains to be evaluated, Furthermore, it will be important to understand whether this interaction differs in the context of the periphery and the brain.

APOE in BBB integrity and function

While APOE exists in two distinct pools in the human body, the influence of each pool on the other, and on AD pathogenesis, remains an important area of research. The most obvious place for the differential influences to converge is at the junction of the periphery and the CNS, the BBB. Studies have shown that the majority of brain Aβ (~ 85%) is cleared across the BBB and the remaining fraction is removed by ISF bulk flow along perivascular spaces [91]. The importance of APOE at the BBB is highlighted by the fact that most cells participating in the formation and maintenance of this complex biological barrier, including astrocytes, pericytes, endothelial cells, smooth muscle cells, and neurons (when stressed), express APOE and/or its receptors [16]. Other cells known to interact with the BBB, such as macrophages and microglia, also express APOE.

It is well known that cerebrovascular damage is a hallmark of AD, as some 80% of all AD patients have cerebral deposition of amyloid, known as cerebral amyloid angiopathy (CAA), which is associated with disruption of the BBB [92]. This dysfunction of the cerebral vasculature has recently re-entered the spotlight in the AD field [9395], and the role of APOE in this process is an area of great interest. Imaging studies show that in cognitively normal APOE4 carriers, there is a widespread and Aβ-independent cerebrovascular dysfunction, indicated by cerebral blood flow (CBF) reduction, impaired cerebrovascular reactivity, and impaired neurovascular coupling [94]. In particular, APOE4 increases the prevalence and severity of CAA [96, 97]. In animal models, early studies indicated that ApoE deficiency leads to damage of the BBB [91, 98100]. More recently, it has been shown that the influence of APOE on BBB integrity, as well as cerebral blood flow, is isoform-dependent [101103]. In APOE4-expressing AD mice, endotoxin-induced peripheral inflammation resulted in reduced cerebral vessel coverage, leakiness of the BBB, and elevated CAA [70]. Further, in a bioengineered 3D model of human vessels, it was shown that while circulating APOE (representing that from the peripheral pool) in general facilitates the clearance of Aβ across the synthetic BBB, APOE4 was deficient in this capability compared to APOE2 [104]. Moreover, efforts are underway to understand the role of peripheral APOE in BBB integrity and function in vivo. It was reported that in mice expressing human APOE3 and APOE4 only by the liver in the mouse ApoE-knockout background, peripheral APOE4, but not APOE3, compromises the BBB integrity and cognitive function [105]. Mechanistically, the role of pericytes in maintaining BBB integrity and clearing vascular Aβ has been underscored recently. APOE2 and APOE3, but not APOE4 or lack of APOE, were shown to suppress the pro-inflammatory signaling cyclophilin A-nuclear factor-κB-matrix-metalloproteinase-9 pathway in pericytes to maintain BBB stability [102]. Similarly, it was found that pericytes actively internalize and degrade Aβ at the BBB, which requires APOE and LRP1; APOE3, but not APOE4, restores Aβ clearance by pericytes when the endogenous ApoE gene is silenced [106]. Thus, pericytes function in an APOE/LRP1 dependent and APOE isoform-specific manner. Together, these findings indicate that amyloid, APOE4 status, and peripheral/central inflammation - all known to induce vascular damage in their own right - may act synergistically to wreak havoc on cerebral blood vessels and the BBB.

APOE in Synaptic Plasticity and Cognitive Function

The role of APOE and its isoforms in synaptic and cognitive function is well documented [107, 108]. Earlier studies have shown that there are APOE isoform-dependent changes in hippocampal synaptic function [109111], and that APOE4-expressing mice exhibit age-dependent cognitive deficits [112, 113], as reported in humans [114, 115]. Recent studies have unraveled some molecular mechanisms underlying these changes associated with APOE4. It was shown that APOE4 selectively impairs APOE receptor trafficking and recycling, leading to reduced glutamate receptor function and synaptic plasticity [116]. Impairment of GABAergic interneurons was identified to cause cognitive deficits in APOE4 mice [113, 117], and enhancing GABA signaling rescues learning and memory function in APOE4 mice [118]. The involvement of GABAergic interneurons in APOE4-induced synaptic dysfunction is further supported by a recent report that APOE4-associated neuronal hyperactivity is caused by reduced responsiveness of excitatory neurons to GABAergic inhibitory inputs [119]. In addition, a recent study demonstrated that APOE4 mice exhibit age-dependent disruption in hippocampal network events critical for memory processes, which depends on GABAergic interneurons [120]. These findings indicate the detrimental effects of APOE4 on synaptic and cognitive function and the importance of the GABAergic system in the maintenance of the excitatory/inhibitory balance in the brain.

The role of peripheral APOE in synaptic/cognitive function has not been extensively studied. As discussed above, APOE4 carriers have low levels of plasma APOE, which may contribute to the increased risk of AD associated with AD. Supporting the connection of peripheral APOE to healthy aging, a recent clinical study of individuals aged 56–105 years showed that centenarians had the highest plasma APOE levels and the lowest frequency of APOE4 allele compared to younger groups [121]. Another study found that peripheral APOE4 was associated with increased loss of grey matter volume in the posterior cingulate, and reduced glucose metabolism in the hippocampus [122]. Furthermore, in cognitively normal E3/E4 carriers, females with higher plasma APOE3 levels performed better on the verbal reasoning (similarities) subtest of the Wechsler Adult Intelligence Scale test. Experimental evidence supporting the importance of plasma APOE came from a recent study, in which expression of APOE only in the periphery improved cognitive performance and partially restored synaptic defects in ApoE-deficient mice [123]. Notably, this study also highlights the importance of APOE in the brain, as synaptic loss and dysfunction still occurred in the brain of ApoE-deficient mice, despite improved memory in the presence of plasma APOE. The role of peripheral APOE in cognitive function is further supported by an elegant recent study, in which ApoE-deficient mice were modified to express either APOE3 or APOEE4 only in the liver; it showed that peripheral APOE3, but not APOE4, enhanced learning and memory performance in the absence of brain APOE [105]. These findings, collectively, indicate that both peripheral and central pools of APOE play important roles in maintaining neurological health.

Emerging therapeutic strategies targeting APOE

As the strongest genetic risk factor for AD, APOE represents the most attractive therapeutic target in AD. However, whether to reduce or increase APOE remains a question of debate, especially with APOE4 exhibiting both loss and gain of functions. In addition, although there is evidence that APOE might be dispensable for cognitive function in the brain, APOE deficiency in the periphery has serious consequences [124126]. In order for an APOE-targeted treatment to be effective, it will be important to understand the differential and cooperative roles that the peripheral and central pools of APOE play in AD pathogenesis, as well as the impact of different APOE isoforms, as discussed in previous sections. Currently pursued therapeutic strategies targeting APOE are summarized in Table 1.

Table 1.

Currently pursued therapeutic strategies targeting APOE

Agent Class Target Proposed Mechanisms Outcomes References
Antibodies APOE Binds Aβ plaqueassociated APOE, specific APOE isoforms (i.e. APOE4), or conformations (i.e. nonlipidated, aggregated) Reduces Aβ plaque deposition and microglial activation, and improves memory in mouse models of AD [127129]
Nuclear receptor agonists LXR, RXR Upregulates APOE and ABCA1 to produce higher levels of APOE, which is more lipidated Reduces Aβ plaque burden and improves memory in mice of all APOE genotypes; conflicting results in human phase II clinical trial [130135]
Apolipoproteinmimetic peptides APOE, APOA-I, APOJ, and their receptors Anti-inflammatory, cardioprotective, improves APOE lipidation, mitigates amyloid toxicity Reduces Aβ plaque burden, protects against neuronal damage, and improves memory in AD mice; safety demonstrated in human phase I clinical trials for some peptides. [139151]
Structure correctors APOE4 Restores the tertiary structure of APOE4 to that of APOE3 Reduces neurotoxicity and Aβ build-up in human stem cells [50, 154]
Genetic approaches (viral vectors) APOE2, APOE4 Converting APOE4 to APOE2, or overexpressing APOE2 in the presence of APOE4 Reduces inflammation and BBB damage, improves APOE lipidation, and reduces Aβ plaque deposition in mice; currently being tested in a phase I human clinical trial. [39, 155]

One exciting approach is passive immunotherapy with APOE-targeted antibodies. Earlier studies showed that antibodies targeting endogenous mouse APOE reduced amyloid accumulation in the brains of mice either before or after the onset of amyloid plaque formation [127, 128]. Recently, it was further shown that antibodies specifically targeting nonlipidated, aggregated human APOE4 inhibited amyloid accumulation in another AD mouse model [129]. Importantly, similar results were obtained with both a central and peripheral delivery of the antibodies, and neither approaches caused significant changes in plasma or brain APOE levels. These findings suggest that the effects of anti-APOE antibodies are not mediated simply by lowering total APOE levels, raising the question of what mechanisms underlying the Aβ-clearing effects of these antibodies. Experimental evidence suggests that microglial-mediated Aβ phagocytosis could be involved in the action of these anti-APOE antibodies [129]. Further studies are needed to unravel the exact mechanisms and to determine any potential adverse effects of this promising approach.

Recognizing the importance of the lipidation state of APOE as well as its overall levels, many groups have focused efforts on drugs targeting the nuclear receptors, e.g., liver X receptor (LXR) and retinoid X receptor (RXR) agonists, as potential therapeutics in AD. These agents have been demonstrated to upregulate ABCA1 and APOE, increase APOE lipidation, and to reduce amyloid pathology and cognitive impairment in mouse models of AD [130]. The most publicized of these agents is bexarotene, owing to its already being a clinically approved agent for the treatment of T-cell lymphoma. Bexarotene was shown to mitigate AD-related pathology and memory deficits in animal models, although some controversy around the effects of this agent on amyloid deposition has arisen [131, 132]. Further complicating the story, a small placebo-controlled phase II clinical trial of bexarotene showed no alteration in amyloid levels in AD patients [133]. However, subgroup analysis determined that APOE4-carriers were holding back the overall analysis, as those carrying APOE2 and APOE3 did in fact have reductions in amyloid burden. This is in contrast to studies in animal models, where bexarotene was shown to mitigate APOE4-driven pathology [134, 135]. In addition, the adverse peripheral effects of bexarotene on triglyceride production and liver health [136138] are of great concern for long-term use of this class of drugs for prevention or treatment of AD.

Alternatively, several groups have investigated the potential of apolipoprotein/high density lipoprotein (HDL) mimetic peptides for modulating APOE lipidation and function. These HDL mimetics possess anti-inflammatory and anti-oxidative properties, and exert vascular-protective effects [11]. A number of these peptides, derived from or based upon key receptor-binding regions of APOA-I, APOE, or APOJ have been shown to improve cognition, reduce neuronal damage, and to mitigate AD-related pathology in animal models [139146]. One particular HDL-mimetic peptide, called 4F, has been tested in three clinical trials for cardiovascular disease and was shown to be safe and well tolerated in humans [147149]. Recent work in our own laboratory with 4F indicates that this unique peptide promotes APOE secretion and lipidation, counteracts Aβ-induced inhibition of APOE secretion and lipidation in primary astrocytes and microglia [150], and reverses APOE4 lipidation deficiency [151].

Another intriguing strategy is based on the observation that differences in APOE protein structure and domain interaction in the nonlipidated or poorly lipidated state underlie the detrimental effects of APOE4, compared with other APOE isoforms [152, 153]. Notably, small-molecule structure-correctors have been designed to restore APOE4 to an APOE3-like tertiary structure, which has been shown to rescue APOE4-associated neuropathology [154]. Recently, this approach was shown to mitigate toxic effect of APOE4 in neurons derived from human iPSCs [50]. Whether these promising molecules affect peripheral and CNS APOE levels and function in vivo awaits further investigation.

Gene therapy is also an important potential therapeutic avenue for targeting APOE. It has been demonstrated that astrocyte-mediated expression of APOE2, delivered intraventricularly via adeno-associated virus (AAV) into the brains of APOE4 mice, improves lipidation and reduces amyloid deposition [39]. This study also demonstrated that injecting APOE4 into APOE4 carriers worsened pathological hallmarks of AD, implying that treatments generally increasing APOE expression may be detrimental in those carrying one or more APOE4 alleles [39]. Importantly, a phase I, open-label clinical trial to test the safety of APOE-targeted gene therapy (NCT03634007) is set to begin recruiting soon. This trial will employ cisternal delivery of an AAVrh.10hAPOE2, serotype rh.10 AAV gene transfer vector expressing human APOE2, into APOE4 homozygotes with positive amyloid PET scans and clinical diagnosis of MCI or AD to determine the maximum tolerated dose, as well as the level and pattern of APOE2 expression in the human brain. This viral vector and route of administration were identified in a previous study to be the safest method for the most widespread delivery of APOE2 in a non-human primate model system [155]. Studies designed to determine the efficacy of this treatment in improving cognitive function in AD will need to wait on the results of this trial. Advances in genetic engineering, such as CRISPR/Cas9, make this an increasingly exciting prospect [51], but much remains to be determined regarding the safety and feasibility of this relatively new technology for prevention or intervention of human disease. Central delivery of any agent is challenging, costly, and potentially dangerous; a fact that is exacerbated by the use of a viral vector, which is likely to induce an inflammatory response as well. An intriguing avenue, supported by many studies discussed in this review, may be to manipulate the APOE genotype in the liver, to influence the make-up of the peripheral pool of APOE.

Limitations of current knowledge

The findings presented here represent significant recent advances in our understanding of APOE biology and its relevance to AD pathogenesis and treatment. Nevertheless, there are limitations to these studies and many knowledge gaps remain. Perhaps the most important limitation herein is our reliance on animal studies, in particular mouse models. While the use of mouse models enables a much wider array of studies than would be possible with human subjects, the relevance to human physiology is not always achieved. As far as lipoprotein metabolism is concerned, one of the major differences between mice and humans is that mice lack cholesteryl ester transfer protein (CETP) [156], which causes HDL being the predominant lipoprotein particles in the circulation of mice. Whereas in humans, the predominant circulating lipoproteins are LDL particles. Thus, mice and humans have very different plasma lipoprotein/cholesterol profiles, consequently different susceptibilities to metabolic and cardiovascular disease. More specifically, human and mouse APOE differ in many respects. Unlike humans, mice only have one form of APOE. Mouse and human APOE only share 70% homology in amino acid sequence [157], and the upstream promoter region of APOE gene shares only 40% homology [158]. Importantly, while both proteins bind the canonical APOE receptors, they also have unique binding partners that are not shared between the homologs. In addition, while APOE is extensively glycosylated in human brain tissue and CSF, mouse brain APOE is scarcely glycosylated [18]. Furthermore, it has been shown that mouse APOE behaves most similarly to human APOE3, as opposed to APOE4, in that it preferentially binds HDL over LDL particles, and does not produce the ‘domain interaction’ thought to underlie some of the detrimental effects of APOE4 [159]. Although human APOE knockin mice have helped addressed some of the primary differences in APOE per se, the inherent lack of CETP in mice remains problematic as a model. Due to the lack of CETP, APOE is necessary for the formation of spherical, lipidated HDL particles in mice; whereas it is not required for genesis of mature HDL particles in humans [160]. Therefore, differences between mouse and human APOE and associated lipoprotein metabolism should be considered when results from mouse models are translated to humans.

Another limitation in our current understanding of APOE is the differential role of therapeutics in the context of APOE genotype. Antibodies targeting APOE2, for instance, may be detrimental, while those targeting APOE4 could produce more protective effects. Similarly, APOE structure corrector molecules would only be hypothesized to provide benefit in APOE4 carriers. On the other hand, Bexarotene and other agents targeting nuclear receptors may be more effective when promoting the expression of APOE2 and APOE3, as opposed to APOE4, as discussed previously. Identifying these differential effects in human carriers of one or more APOE genotype will be critical for the future success of APOE-targeted therapeutics.

Concluding remarks

The current evidence highlights the far-reaching impact of APOE in AD. From peripheral tissues to different cell types in the brain, APOE plays a pivotal role, and APOE4 drives pathology (Fig. 2). With a continued need for effective therapeutics to halt the clinical progression of AD, and setbacks in recent clinical trials, APOE has rapidly risen to the forefront as a druggable target. It will be important to understand how APOE-targeted therapeutics modify potential comorbidities, including cardiovascular disease and systemic inflammation, as well as any potential adverse events associated with systemic administration of these agents. While central pools of APOE drive key pathological hallmarks of AD, including amyloid deposition, tau hyperphosphorylation, and neuroinflammation, peripheral APOE contributes to systemic inflammation, amyloid clearance, BBB integrity, and overall vascular health. In both the periphery and in the brain, APOE4 is a key driver of pathological progression. Many questions remain unanswered as to the direct and indirect impact of peripheral APOE on cognition and AD, and a focus on this area will be vital to continued efforts to combat this horrendous disease.

Fig. 2.

Fig. 2.

APOE4 in the periphery and in the brain drives the pathological progression to AD.

Highlights.

  • APOE exists in two distinct pools, in the periphery and in the brain.

  • APOE regulates multiple molecular pathways in addition to cholesterol homeostasis.

  • The three APOE isoforms exert differential effects on Alzheimer-related processes.

  • Both peripheral and central pools of APOE contribute to Alzheimer’s disease risk.

  • Promising APOE-targeted therapeutic strategies have emerged for Alzheimer’s disease.

Acknowledgements

This work was supported in part by the National Institutes of Health [grant numbers AG056025, AG058081, and AG056976]. DC was supported by a predoctoral training fellowship in the PharmacoNeuroImmunology Program from the National Institutes of Health (grant number DA007097) and a 3M Science and Technology Training Fellowship. AG is partly supported by the Kwanjeong Educational Foundation Overseas Scholarship from South Korea.

Footnotes

Disclosure statement

The authors report no declarations of interest

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • [1].Mahley RW, Apolipoprotein E: cholesterol transport protein with expanding role in cell biology, Science 240 (1988) 622–630. [DOI] [PubMed] [Google Scholar]
  • [2].Mahley RW, Central Nervous System Lipoproteins: ApoE and Regulation of Cholesterol Metabolism, Arterioscler Thromb Vasc Biol 36 (2016) 1305–1315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [3].Holtzman DM, Herz J, Bu G, Apolipoprotein E and apolipoprotein E receptors: normal biology and roles in Alzheimer disease, Cold Spring Harbor perspectives in medicine 2 (2012) a006312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [4].Zhao N, Liu CC, Qiao W, Bu G, Apolipoprotein E, Receptors, and Modulation of Alzheimer’s Disease, Biol Psychiatry 83 (2018) 347–357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [5].Weisgraber KH, Rall SC Jr., Mahley RW, Human E apoprotein heterogeneity. Cysteine-arginine interchanges in the amino acid sequence of the apo-E isoforms, J Biol Chem 256 (1981) 9077–9083. [PubMed] [Google Scholar]
  • [6].Rall SC Jr., Weisgraber KH, Mahley RW, Human apolipoprotein E. The complete amino acid sequence, J Biol Chem 257 (1982) 4171–4178. [PubMed] [Google Scholar]
  • [7].Mahley RW, Apolipoprotein E: from cardiovascular disease to neurodegenerative disorders, Journal of molecular medicine 94 (2016) 739–746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [8].Corder EH, Saunders AM, Strittmatter WJ, Schmechel DE, Gaskell PC, Small GW, Roses AD, Haines JL, Pericak-Vance MA, Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer’s disease in late onset families, Science 261 (1993) 921–923. [DOI] [PubMed] [Google Scholar]
  • [9].Saunders AM, Strittmatter WJ, Schmechel D, George-Hyslop PH, Pericak-Vance MA, Joo SH, Rosi BL, Gusella JF, Crapper-MacLachlan DR, Alberts MJ, et al. , Association of apolipoprotein E allele epsilon 4 with late-onset familial and sporadic Alzheimer’s disease, Neurology 43 (1993) 1467–1472. [DOI] [PubMed] [Google Scholar]
  • [10].Strittmatter WJ, Saunders AM, Schmechel D, Pericak-Vance M, Enghild J, Salvesen GS, Roses AD, Apolipoprotein E: high-avidity binding to beta-amyloid and increased frequency of type 4 allele in late-onset familial Alzheimer disease, Proc Natl Acad Sci U S A 90 (1993) 1977–1981. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [11].Hottman DA, Chernick D, Cheng S, Wang Z, Li L, HDL and cognition in neurodegenerative disorders, Neurobiol Dis 72 Pt A (2014) 22–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [12].Vitali C, Wellington CL, Calabresi L, HDL and cholesterol handling in the brain, Cardiovasc Res 103 (2014) 405–413. [DOI] [PubMed] [Google Scholar]
  • [13].Linton MF, Gish R, Hubl ST, Butler E, Esquivel C, Bry WI, Boyles JK, Wardell MR, Young SG, Phenotypes of apolipoprotein B and apolipoprotein E after liver transplantation, J Clin Invest 88 (1991) 270–281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [14].Xu Q, Bernardo A, Walker D, Kanegawa T, Mahley RW, Huang Y, Profile and regulation of apolipoprotein E (ApoE) expression in the CNS in mice with targeting of green fluorescent protein gene to the ApoE locus, J Neurosci 26 (2006) 4985–4994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [15].Zhang Y, Chen K, Sloan SA, Bennett ML, Scholze AR, O’Keeffe S, Phatnani HP, Guarnieri P, Caneda C, Ruderisch N, Deng S, Liddelow SA, Zhang C, Daneman R, Maniatis T, Barres BA, Wu JQ, An RNA-sequencing transcriptome and splicing database of glia, neurons, and vascular cells of the cerebral cortex, J Neurosci 34 (2014) 11929–11947. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Kanekiyo T, Xu H, Bu G, ApoE and Abeta in Alzheimer’s Disease: Accidental Encounters or Partners?, Neuron 81 (2014) 740–754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Braesch-Andersen S, Paulie S, Smedman C, Mia S, Kumagai-Braesch M, ApoE production in human monocytes and its regulation by inflammatory cytokines, PLoS One 8 (2013) e79908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [18].Kockx M, Traini M, Kritharides L, Cell-specific production, secretion, and function of apolipoprotein E, Journal of molecular medicine 96 (2018) 361–371. [DOI] [PubMed] [Google Scholar]
  • [19].Koch S, Donarski N, Goetze K, Kreckel M, Stuerenburg HJ, Buhmann C, Beisiegel U, Characterization of four lipoprotein classes in human cerebrospinal fluid, J Lipid Res 42 (2001) 1143–1151. [PubMed] [Google Scholar]
  • [20].Panza F, Solfrizzi V, Colacicco AM, Basile AM, 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 14 (2003) 605–608. [DOI] [PubMed] [Google Scholar]
  • [21].Boisvert WA, Spangenberg J, Curtiss LK, Treatment of severe hypercholesterolemia in apolipoprotein E-deficient mice by bone marrow transplantation, J Clin Invest 96 (1995) 1118–1124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [22].Linton MF, Atkinson JB, Fazio S, Prevention of atherosclerosis in apolipoprotein E-deficient mice by bone marrow transplantation, Science 267 (1995) 1034–1037. [DOI] [PubMed] [Google Scholar]
  • [23].Xu Q, Walker D, Bernardo A, Brodbeck J, Balestra ME, Huang Y, Intron-3 retention/splicing controls neuronal expression of apolipoprotein E in the CNS, J Neurosci 28 (2008) 1452–1459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [24].Kockx M, Jessup W, Kritharides L, Regulation of endogenous apolipoprotein E secretion by macrophages, Arterioscler Thromb Vasc Biol 28 (2008) 1060–1067. [DOI] [PubMed] [Google Scholar]
  • [25].DiBattista AM, Dumanis SB, Newman J, Rebeck GW, Identification and modification of amyloid-independent phenotypes of APOE4 mice, Exp Neurol 280 (2016) 97–105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [26].Cruchaga C, Kauwe JS, Nowotny P, Bales K, Pickering EH, Mayo K, Bertelsen S, Hinrichs A, Alzheimer’s Disease Neuroimaging I, Fagan AM, Holtzman DM, Morris JC, Goate AM, Cerebrospinal fluid APOE levels: an endophenotype for genetic studies for Alzheimer’s disease, Hum Mol Genet 21 (2012) 4558–4571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [27].Martinez-Morillo E, Hansson O, Atagi Y, Bu G, Minthon L, Diamandis EP, Nielsen HM, Total apolipoprotein E levels and specific isoform composition in cerebrospinal fluid and plasma from Alzheimer’s disease patients and controls, Acta Neuropathol 127 (2014) 633–643. [DOI] [PubMed] [Google Scholar]
  • [28].Baker-Nigh AT, Mawuenyega KG, Bollinger JG, Ovod V, Kasten T, Franklin EE, Liao F, Jiang H, Holtzman D, Cairns NJ, Morris JC, Bateman RJ, Human Central Nervous System (CNS) ApoE Isoforms Are Increased by Age, Differentially Altered by Amyloidosis, and Relative Amounts Reversed in the CNS Compared with Plasma, J Biol Chem 291 (2016) 27204–27218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [29].Rasmussen KL, Tybjaerg-Hansen A, Nordestgaard BG, Frikke-Schmidt R, Plasma levels of apolipoprotein E and risk of dementia in the general population, Ann Neurol 77 (2015) 301–311. [DOI] [PubMed] [Google Scholar]
  • [30].Wolters FJ, Koudstaal PJ, Hofman A, van Duijn CM, Ikram MA, Serum apolipoprotein E is associated with long-term risk of Alzheimer’s disease: The Rotterdam Study, Neuroscience Letters 617 (2016) 139–142. [DOI] [PubMed] [Google Scholar]
  • [31].Rasmussen KL, Tybjaerg-Hansen A, Nordestgaard BG, Frikke-Schmidt R, Plasma apolipoprotein E levels and risk of dementia: A Mendelian randomization study of 106,562 individuals, Alzheimers Dement 14 (2018) 71–80. [DOI] [PubMed] [Google Scholar]
  • [32].Zhang SH, Reddick RL, Piedrahita JA, Maeda N, Spontaneous hypercholesterolemia and arterial lesions in mice lacking apolipoprotein E, Science 258 (1992) 468–471. [DOI] [PubMed] [Google Scholar]
  • [33].Plump AS, Smith JD, Hayek T, Aalto-Setala K, Walsh A, Verstuyft JG, Rubin EM, Breslow JL, Severe hypercholesterolemia and atherosclerosis in apolipoprotein E-deficient mice created by homologous recombination in ES cells, Cell 71 (1992) 343–353. [DOI] [PubMed] [Google Scholar]
  • [34].Dietschy JM, Turley SD, Cholesterol metabolism in the brain, Curr Opin Lipidol 12 (2001) 105–112. [DOI] [PubMed] [Google Scholar]
  • [35].Dietschy JM, Turley SD, Thematic review series: brain Lipids. Cholesterol metabolism in the central nervous system during early development and in the mature animal, J Lipid Res 45 (2004) 1375–1397. [DOI] [PubMed] [Google Scholar]
  • [36].Michikawa M, Fan QW, Isobe I, Yanagisawa K, Apolipoprotein E exhibits isoform-specific promotion of lipid efflux from astrocytes and neurons in culture, J Neurochem 74 (2000) 1008–1016. [DOI] [PubMed] [Google Scholar]
  • [37].Gong JS, Kobayashi M, Hayashi H, Zou K, Sawamura N, Fujita SC, Yanagisawa K, Michikawa M, Apolipoprotein E (ApoE) isoform-dependent lipid release from astrocytes prepared from human ApoE3 and ApoE4 knock-in mice, J Biol Chem 277 (2002) 29919–29926. [DOI] [PubMed] [Google Scholar]
  • [38].Tai LM, Mehra S, Shete V, Estus S, Rebeck GW, Bu G, LaDu MJ, Soluble apoE/Abeta complex: mechanism and therapeutic target for APOE4-induced AD risk, Mol Neurodegener 9 (2014) 2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [39].Hu J, Liu CC, Chen XF, Zhang YW, Xu H, Bu G, Opposing effects of viral mediated brain expression of apolipoprotein E2 (apoE2) and apoE4 on apoE lipidation and Abeta metabolism in apoE4-targeted replacement mice, Mol Neurodegener 10 (2015) 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [40].Hanson AJ, Bayer-Carter JL, Green PS, Montine TJ, Wilkinson CW, Baker LD, Watson GS, Bonner LM, Callaghan M, Leverenz JB, Tsai E, Postupna N, Zhang J, Lampe J, Craft S, Effect of apolipoprotein E genotype and diet on apolipoprotein E lipidation and amyloid peptides: randomized clinical trial, JAMA neurology 70 (2013) 972–980. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [41].Heinsinger NM, Gachechiladze MA, Rebeck GW, Apolipoprotein E Genotype Affects Size of ApoE Complexes in Cerebrospinal Fluid, J Neuropathol Exp Neurol 75 (2016) 918–924. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [42].Zhu L, Zhong M, Elder GA, Sano M, Holtzman DM, Gandy S, Cardozo C, Haroutunian V, Robakis NK, Cai D, Phospholipid dysregulation contributes to ApoE4-associated cognitive deficits in Alzheimer’s disease pathogenesis, Proc Natl Acad Sci U S A 112 (2015) 11965–11970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [43].Huang Y, Mahley RW, Apolipoprotein E: structure and function in lipid metabolism, neurobiology, and Alzheimer’s diseases, Neurobiol Dis 72 Pt A (2014) 3–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [44].Liao F, Yoon H, Kim J, Apolipoprotein E metabolism and functions in brain and its role in Alzheimer’s disease, Curr Opin Lipidol 28 (2017) 60–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [45].Kim J, Basak JM, Holtzman DM, The role of apolipoprotein E in Alzheimer’s disease, Neuron 63 (2009) 287–303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [46].Liu CC, Zhao N, Fu Y, Wang N, Linares C, Tsai CW, Bu G, ApoE4 Accelerates Early Seeding of Amyloid Pathology, Neuron 96 (2017) 1024–1032 e1023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [47].Huynh TV, Liao F, Francis CM, Robinson GO, Serrano JR, Jiang H, Roh J, Finn MB, Sullivan PM, Esparza TJ, Stewart FR, Mahan TE, Ulrich JD, Cole T, Holtzman DM, Age-Dependent Effects of apoE Reduction Using Antisense Oligonucleotides in a Model of beta-amyloidosis, Neuron 96 (2017) 1013–1023 e1014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [48].Habchi J, Chia S, Galvagnion C, Michaels TCT, Bellaiche MMJ, Ruggeri FS, Sanguanini M, Idini I, Kumita JR, Sparr E, Linse S, Dobson CM, Knowles TPJ, Vendruscolo M, Cholesterol catalyses Abeta42 aggregation through a heterogeneous nucleation pathway in the presence of lipid membranes, Nature chemistry 10 (2018) 673–683. [DOI] [PubMed] [Google Scholar]
  • [49].Huang YA, Zhou B, Wernig M, Sudhof TC, ApoE2, ApoE3, and ApoE4 Differentially Stimulate APP Transcription and Abeta Secretion, Cell 168 (2017) 427–441 e421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [50].Wang C, Najm R, Xu Q, Jeong DE, Walker D, Balestra ME, Yoon SY, Yuan H, Li G, Miller ZA, Miller BL, Malloy MJ, Huang Y, Gain of toxic apolipoprotein E4 effects in human iPSC-derived neurons is ameliorated by a small-molecule structure corrector, Nat Med 24 (2018) 647–657. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [51].Lin YT, Seo J, Gao F, Feldman HM, Wen HL, Penney J, Cam HP, Gjoneska E, Raja WK, Cheng J, Rueda R, Kritskiy O, Abdurrob F, Peng Z, Milo B, Yu CJ, Elmsaouri S, Dey D, Ko T, Yankner BA, Tsai LH, APOE4 Causes Widespread Molecular and Cellular Alterations Associated with Alzheimer’s Disease Phenotypes in Human iPSC-Derived Brain Cell Types, Neuron 98 (2018) 1141–1154 e1147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [52].Shi Y, Yamada K, Liddelow SA, Smith ST, Zhao L, Luo W, Tsai RM, Spina S, Grinberg LT, Rojas JC, Gallardo G, Wang K, Roh J, Robinson G, Finn MB, Jiang H, Sullivan PM, Baufeld C, Wood MW, Sutphen C, McCue L, Xiong C, Del-Aguila JL, Morris JC, Cruchaga C, Alzheimer’s Disease Neuroimaging I, Fagan AM, Miller BL, Boxer AL, Seeley WW, Butovsky O, Barres BA, Paul SM, Holtzman DM, ApoE4 markedly exacerbates tau-mediated neurodegeneration in a mouse model of tauopathy, Nature 549 (2017) 523–527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [53].Cao J, Gaamouch FE, Meabon JS, Meeker KD, Zhu L, Zhong MB, Bendik J, Elder G, Jing P, Xia J, Luo W, Cook DG, Cai D, ApoE4-associated phospholipid dysregulation contributes to development of Tau hyper-phosphorylation after traumatic brain injury, Scientific reports 7 (2017) 11372. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [54].Badia MC, Lloret A, Giraldo E, Dasi F, Olaso G, Alonso MD, Vina J, Lymphocytes from young healthy persons carrying the ApoE4 allele overexpress stress-related proteins involved in the pathophysiology of Alzheimer’s disease, J Alzheimers Dis 33 (2013) 77–83. [DOI] [PubMed] [Google Scholar]
  • [55].Zhao N, Liu CC, Van Ingelgom AJ, Linares C, Kurti A, Knight JA, Heckman MG, Diehl NN, Shinohara M, Martens YA, Attrebi ON, Petrucelli L, Fryer JD, Wszolek ZK, Graff-Radford NR, Caselli RJ, Sanchez-Contreras MY, Rademakers R, Murray ME, Koga S, Dickson DW, Ross OA, Bu G, APOE epsilon2 is associated with increased tau pathology in primary tauopathy, Nature communications 9 (2018) 4388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [56].Farfel JM, Yu L, De Jager PL, Schneider JA, Bennett DA, Association of APOE with tau-tangle pathology with and without beta-amyloid, Neurobiol Aging 37 (2016) 19–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [57].Barger SW, Harmon AD, Microglial activation by Alzheimer amyloid precursor protein and modulation by apolipoprotein E, Nature 388 (1997) 878–881. [DOI] [PubMed] [Google Scholar]
  • [58].Lynch JR, Morgan D, Mance J, Matthew WD, Laskowitz DT, Apolipoprotein E modulates glial activation and the endogenous central nervous system inflammatory response, J Neuroimmunol 114 (2001) 107–113. [DOI] [PubMed] [Google Scholar]
  • [59].Vitek MP, Brown CM, Colton CA, APOE genotype-specific differences in the innate immune response, Neurobiol Aging 30 (2009) 1350–1360. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [60].Gale SC, Gao L, Mikacenic C, Coyle SM, Rafaels N, Murray Dudenkov T, Madenspacher JH, Draper DW, Ge W, Aloor JJ, Azzam KM, Lai L, Blackshear PJ, Calvano SE, Barnes KC, Lowry SF, Corbett S, Wurfel MM, Fessler MB, APOepsilon4 is associated with enhanced in vivo innate immune responses in human subjects, The Journal of allergy and clinical immunology 134 (2014) 127–134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [61].Desikan RS, Schork AJ, Wang Y, Thompson WK, Dehghan A, Ridker PM, Chasman DI, McEvoy LK, Holland D, Chen CH, Karow DS, Brewer JB, Hess CP, Williams J, Sims R, O’Donovan MC, Choi SH, Bis JC, Ikram MA, Gudnason V, DeStefano AL, van der Lee SJ, Psaty BM, van Duijn CM, Launer L, Seshadri S, Pericak-Vance MA, Mayeux R, Haines JL, Farrer LA, Hardy J, Ulstein ID, Aarsland D, Fladby T, White LR, Sando SB, Rongve A, Witoelar A, Djurovic S, Hyman BT, Snaedal J, Steinberg S, Stefansson H, Stefansson K, Schellenberg GD, Andreassen OA, Dale AM, Inflammation Working G., International P Genomics of Alzheimer’s Disease, I. DemGene, Polygenic Overlap Between C-Reactive Protein, Plasma Lipids, and Alzheimer Disease, Circulation 131 (2015) 2061–2069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [62].Tan ZS, Beiser AS, Vasan RS, Roubenoff R, Dinarello CA, Harris TB, Benjamin EJ, Au R, Kiel DP, Wolf PA, Seshadri S, Inflammatory markers and the risk of Alzheimer disease: the Framingham Study, Neurology 68 (2007) 1902–1908. [DOI] [PubMed] [Google Scholar]
  • [63].O’Bryant SE, Waring SC, Hobson V, Hall JR, Moore CB, Bottiglieri T, Massman P, Diaz-Arrastia R, Decreased C-reactive protein levels in Alzheimer disease, J Geriatr Psychiatry Neurol 23 (2010) 49–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [64].Hu WT, Holtzman DM, Fagan AM, Shaw LM, Perrin R, Arnold SE, Grossman M, Xiong C, Craig-Schapiro R, Clark CM, Pickering E, Kuhn M, Chen Y, Van Deerlin VM, McCluskey L, Elman L, Karlawish J, Chen-Plotkin A, Hurtig HI, Siderowf A, Swenson F, Lee VM, Morris JC, Trojanowski JQ, Soares H, Alzheimer’s Disease Neuroimaging I, Plasma multianalyte profiling in mild cognitive impairment and Alzheimer disease, Neurology 79 (2012) 897–905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [65].Schmidt R, Schmidt H, Curb JD, Masaki K, White LR, Launer LJ, Early inflammation and dementia: a 25-year follow-up of the Honolulu-Asia Aging Study, Ann Neurol 52 (2002) 168–174. [DOI] [PubMed] [Google Scholar]
  • [66].Engelhart MJ, Geerlings MI, Meijer J, Kiliaan A, Ruitenberg A, van Swieten JC, Stijnen T, Hofman A, Witteman JC, Breteler MM, Inflammatory proteins in plasma and the risk of dementia: the rotterdam study, Arch Neurol 61 (2004) 668–672. [DOI] [PubMed] [Google Scholar]
  • [67].Schuitemaker A, Dik MG, Veerhuis R, Scheltens P, Schoonenboom NS, Hack CE, Blankenstein MA, Jonker C, Inflammatory markers in AD and MCI patients with different biomarker profiles, Neurobiol Aging 30 (2009) 1885–1889. [DOI] [PubMed] [Google Scholar]
  • [68].Tao Q, Ang TFA, DeCarli C, Auerbach SH, Devine S, Stein TD, Zhang X, Massaro J, Au R, Qiu WQ, Association of Chronic Low-grade Inflammation With Risk of Alzheimer Disease in ApoE4 Carriers, JAMA Netw Open 1 (2018) e183597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [69].Royall DR, Al-Rubaye S, Bishnoi R, Palmer RF, Few serum proteins mediate APOE’s association with dementia, PLoS One 12 (2017) e0172268. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [70].Marottoli FM, Katsumata Y, Koster KP, Thomas R, Fardo DW, Tai LM, Peripheral Inflammation, Apolipoprotein E4, and Amyloid-beta Interact to Induce Cognitive and Cerebrovascular Dysfunction, ASN neuro 9 (2017) 1759091417719201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [71].Teter B, LaDu MJ, Sullivan PM, Frautschy SA, Cole GM, Apolipoprotein E isotype-dependent modulation of microRNA-146a in plasma and brain, Neuroreport 27 (2016) 791–795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [72].Guerreiro R, Wojtas A, Bras J, Carrasquillo M, Rogaeva E, Majounie E, Cruchaga C, Sassi C, Kauwe JS, Younkin S, Hazrati L, Collinge J, Pocock J, Lashley T, Williams J, Lambert JC, Amouyel P, Goate A, Rademakers R, Morgan K, Powell J, St George-Hyslop P, Singleton A, Hardy J, Alzheimer Genetic Analysis G, TREM2 variants in Alzheimer’s disease, N Engl J Med 368 (2013) 117–127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [73].Jonsson T, Stefansson H, Steinberg S, Jonsdottir I, Jonsson PV, Snaedal J, Bjornsson S, Huttenlocher J, Levey AI, Lah JJ, Rujescu D, Hampel H, Giegling I, Andreassen OA, Engedal K, Ulstein I, Djurovic S, Ibrahim-Verbaas C, Hofman A, Ikram MA, van Duijn CM, Thorsteinsdottir U, Kong A, Stefansson K, Variant of TREM2 associated with the risk of Alzheimer’s disease, N Engl J Med 368 (2013) 107–116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [74].Colonna M, Wang Y, TREM2 variants: new keys to decipher Alzheimer disease pathogenesis, Nat Rev Neurosci 17 (2016) 201–207. [DOI] [PubMed] [Google Scholar]
  • [75].Ulrich JD, Ulland TK, Colonna M, Holtzman DM, Elucidating the Role of TREM2 in Alzheimer’s Disease, Neuron 94 (2017) 237–248. [DOI] [PubMed] [Google Scholar]
  • [76].Yeh FL, Hansen DV, Sheng M, TREM2, Microglia, and Neurodegenerative Diseases, Trends Mol Med 23 (2017) 512–533. [DOI] [PubMed] [Google Scholar]
  • [77].Jay TR, von Saucken VE, Landreth GE, TREM2 in Neurodegenerative Diseases, Mol Neurodegener 12 (2017) 56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [78].Atagi Y, Liu CC, Painter MM, Chen XF, Verbeeck C, Zheng H, Li X, Rademakers R, Kang SS, Xu H, Younkin S, Das P, Fryer JD, Bu G, Apolipoprotein E Is a Ligand for Triggering Receptor Expressed on Myeloid Cells 2 (TREM2), J Biol Chem 290 (2015) 26043–26050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [79].Bailey CC, DeVaux LB, Farzan M, The Triggering Receptor Expressed on Myeloid Cells 2 Binds Apolipoprotein E, J Biol Chem 290 (2015) 26033–26042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [80].Krasemann S, Madore C, Cialic R, Baufeld C, Calcagno N, El Fatimy R, Beckers L, O’Loughlin E, Xu Y, Fanek Z, Greco DJ, Smith ST, Tweet G, Humulock Z, Zrzavy T, Conde-Sanroman P, Gacias M, Weng Z, Chen H, Tjon E, Mazaheri F, Hartmann K, Madi A, Ulrich JD, Glatzel M, Worthmann A, Heeren J, Budnik B, Lemere C, Ikezu T, Heppner FL, Litvak V, Holtzman DM, Lassmann H, Weiner HL, Ochando J, Haass C, Butovsky O, The TREM2-APOE Pathway Drives the Transcriptional Phenotype of Dysfunctional Microglia in Neurodegenerative Diseases, Immunity 47 (2017) 566–581 e569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [81].Yeh FL, Wang Y, Tom I, Gonzalez LC, Sheng M, TREM2 Binds to Apolipoproteins, Including APOE and CLU/APOJ, and Thereby Facilitates Uptake of Amyloid-Beta by Microglia, Neuron 91 (2016) 328–340. [DOI] [PubMed] [Google Scholar]
  • [82].Jendresen C, Arskog V, Daws MR, Nilsson LN, The Alzheimer’s disease risk factors apolipoprotein E and TREM2 are linked in a receptor signaling pathway, J Neuroinflammation 14 (2017) 59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [83].Keren-Shaul H, Spinrad A, Weiner A, Matcovitch-Natan O, Dvir-Szternfeld R, Ulland TK, David E, Baruch K, Lara-Astaiso D, Toth B, Itzkovitz S, Colonna M, Schwartz M, Amit I, A Unique Microglia Type Associated with Restricting Development of Alzheimer’s Disease, Cell 169 (2017) 1276–1290 e1217. [DOI] [PubMed] [Google Scholar]
  • [84].Kang SS, Ebbert MTW, Baker KE, Cook C, Wang X, Sens JP, Kocher JP, Petrucelli L, Fryer JD, Microglial translational profiling reveals a convergent APOE pathway from aging, amyloid, and tau, J Exp Med 215 (2018) 2235–2245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [85].Parhizkar S, Arzberger T, Brendel M, Kleinberger G, Deussing M, Focke C, Nuscher B, Xiong M, Ghasemigharagoz A, Katzmarski N, Krasemann S, Lichtenthaler SF, Muller SA, Colombo A, Monasor LS, Tahirovic S, Herms J, Willem M, Pettkus N, Butovsky O, Bartenstein P, Edbauer D, Rominger A, Erturk A, Grathwohl SA, Neher JJ, Holtzman DM, Meyer-Luehmann M, Haass C, Loss of TREM2 function increases amyloid seeding but reduces plaque-associated ApoE, Nat Neurosci 22 (2019) 191–204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [86].Ulrich JD, Ulland TK, Mahan TE, Nystrom S, Nilsson KP, Song WM, Zhou Y, Reinartz M, Choi S, Jiang H, Stewart FR, Anderson E, Wang Y, Colonna M, Holtzman DM, ApoE facilitates the microglial response to amyloid plaque pathology, J Exp Med 215 (2018) 1047–1058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [87].Shi Y, Holtzman DM, Interplay between innate immunity and Alzheimer disease: APOE and TREM2 in the spotlight, Nature reviews. Immunology (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [88].Kober DL, Brett TJ, TREM2-Ligand Interactions in Health and Disease, J Mol Biol 429 (2017) 1607–1629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [89].Hu N, Tan MS, Yu JT, Sun L, Tan L, Wang YL, Jiang T, Tan L, Increased expression of TREM2 in peripheral blood of Alzheimer’s disease patients, J Alzheimers Dis 38 (2014) 497–501. [DOI] [PubMed] [Google Scholar]
  • [90].Casati M, Ferri E, Gussago C, Mazzola P, Abbate C, Bellelli G, Mari D, Cesari M, Arosio B, Increased expression of TREM2 in peripheral cells from mild cognitive impairment patients who progress into Alzheimer’s disease, European journal of neurology : the official journal of the European Federation of Neurological Societies 25 (2018) 805–810. [DOI] [PubMed] [Google Scholar]
  • [91].Shibata M, Yamada S, Kumar SR, Calero M, Bading J, Frangione B, Holtzman DM, Miller CA, Strickland DK, Ghiso J, Zlokovic BV, Clearance of Alzheimer’s amyloid-ss(1–40) peptide from brain by LDL receptor-related protein-1 at the blood-brain barrier, J Clin Invest 106 (2000) 1489–1499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [92].Serrano-Pozo A, Frosch MP, Masliah E, Hyman BT, Neuropathological alterations in Alzheimer disease, Cold Spring Harbor perspectives in medicine 1 (2011) a006189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [93].Montagne A, Zhao Z, Zlokovic BV, Alzheimer’s disease: A matter of blood-brain barrier dysfunction?, J Exp Med 214 (2017) 3151–3169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [94].Sweeney MD, Kisler K, Montagne A, Toga AW, Zlokovic BV, The role of brain vasculature in neurodegenerative disorders, Nat Neurosci 21 (2018) 1318–1331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [95].Sweeney MD, Sagare AP, Zlokovic BV, Blood-brain barrier breakdown in Alzheimer disease and other neurodegenerative disorders, Nature reviews. Neurology 14 (2018) 133–150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [96].Ringman JM, Sachs MC, Zhou Y, Monsell SE, Saver JL, Vinters HV, Clinical predictors of severe cerebral amyloid angiopathy and influence of APOE genotype in persons with pathologically verified Alzheimer disease, JAMA neurology 71 (2014) 878–883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [97].Esiri M, Chance S, Joachim C, Warden D, Smallwood A, Sloan C, Christie S, Wilcock G, Smith AD, Cerebral amyloid angiopathy, subcortical white matter disease and dementia: literature review and study in OPTIMA, Brain Pathol 25 (2015) 51–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [98].Hafezi-Moghadam A, Thomas KL, Wagner DD, ApoE deficiency leads to a progressive age-dependent blood-brain barrier leakage, Am J Physiol Cell Physiol 292 (2007) C1256–1262. [DOI] [PubMed] [Google Scholar]
  • [99].Methia N, Andre P, Hafezi-Moghadam A, Economopoulos M, Thomas KL, Wagner DD, ApoE deficiency compromises the blood brain barrier especially after injury, Mol Med 7 (2001) 810–815. [PMC free article] [PubMed] [Google Scholar]
  • [100].Fullerton SM, Shirman GA, Strittmatter WJ, Matthew WD, Impairment of the blood-nerve and blood-brain barriers in apolipoprotein e knockout mice, Exp Neurol 169 (2001) 13–22. [DOI] [PubMed] [Google Scholar]
  • [101].Nishitsuji K, Hosono T, Nakamura T, Bu G, Michikawa M, Apolipoprotein E regulates the integrity of tight junctions in an isoform-dependent manner in an in vitro blood-brain barrier model, J Biol Chem 286 (2011) 17536–17542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [102].Bell RD, Winkler EA, Singh I, Sagare AP, Deane R, Wu Z, Holtzman DM, Betsholtz C, Armulik A, Sallstrom J, Berk BC, Zlokovic BV, Apolipoprotein E controls cerebrovascular integrity via cyclophilin A, Nature 485 (2012) 512–516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [103].Alata W, Ye Y, St-Amour I, Vandal M, Calon F, Human apolipoprotein E varepsilon4 expression impairs cerebral vascularization and blood-brain barrier function in mice, J Cereb Blood Flow Metab 35 (2015) 86–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [104].Robert J, Button EB, Yuen B, Gilmour M, Kang K, Bahrabadi A, Stukas S, Zhao W, Kulic I, Wellington CL, Clearance of beta-amyloid is facilitated by apolipoprotein E and circulating high-density lipoproteins in bioengineered human vessels, eLife 6 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [105].Shugart D, ApoE Has Hand in Alzheimer’s Beyond Aβ, Beyond the Brain, https://www.alzforum.org/news/research-news/apoe-has-hand-alzheimers-beyond-av-beyond-brain (2018).
  • [106].Ma Q, Zhao Z, Sagare AP, Wu Y, Wang M, Owens NC, Verghese PB, Herz J, Holtzman DM, Zlokovic BV, Blood-brain barrier-associated pericytes internalize and clear aggregated amyloid-beta42 by LRP1-dependent apolipoprotein E isoform-specific mechanism, Mol Neurodegener 13 (2018) 57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [107].Kim J, Yoon H, Basak J, Kim J, Apolipoprotein E in synaptic plasticity and Alzheimer’s disease: potential cellular and molecular mechanisms, Molecules and cells 37 (2014) 767–776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [108].Di Battista AM, Heinsinger NM, Rebeck GW, Alzheimer’s Disease Genetic Risk Factor APOE-epsilon4 Also Affects Normal Brain Function, Curr Alzheimer Res 13 (2016) 1200–1207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [109].Trommer BL, Shah C, Yun SH, Gamkrelidze G, Pasternak ES, Ye GL, Sotak M, Sullivan PM, Pasternak JF, LaDu MJ, ApoE isoform affects LTP in human targeted replacement mice, Neuroreport 15 (2004) 2655–2658. [DOI] [PubMed] [Google Scholar]
  • [110].Trommer BL, Shah C, Yun SH, Gamkrelidze G, Pasternak ES, Stine WB, Manelli A, Sullivan P, Pasternak JF, LaDu MJ, ApoE isoform-specific effects on LTP: blockade by oligomeric amyloid-beta1–42, Neurobiol Dis 18 (2005) 75–82. [DOI] [PubMed] [Google Scholar]
  • [111].Korwek KM, Trotter JH, Ladu MJ, Sullivan PM, Weeber EJ, ApoE isoform-dependent changes in hippocampal synaptic function, Mol Neurodegener 4 (2009) 21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [112].Raber J, Wong D, Yu GQ, Buttini M, Mahley RW, Pitas RE, Mucke L, Apolipoprotein E and cognitive performance, Nature 404 (2000) 352–354. [DOI] [PubMed] [Google Scholar]
  • [113].Andrews-Zwilling Y, Bien-Ly N, Xu Q, Li G, Bernardo A, Yoon SY, Zwilling D, Yan TX, Chen L, Huang Y, Apolipoprotein E4 causes age- and Tau-dependent impairment of GABAergic interneurons, leading to learning and memory deficits in mice, J Neurosci 30 (2010) 13707–13717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [114].Mayeux R, Small SA, Tang M, Tycko B, Stern Y, Memory performance in healthy elderly without Alzheimer’s disease: effects of time and apolipoprotein-E, Neurobiol Aging 22 (2001) 683–689. [DOI] [PubMed] [Google Scholar]
  • [115].Dik MG, Jonker C, Comijs HC, Bouter LM, Twisk JW, van Kamp GJ, Deeg DJ, Memory complaints and APOE-epsilon4 accelerate cognitive decline in cognitively normal elderly, Neurology 57 (2001) 2217–2222. [DOI] [PubMed] [Google Scholar]
  • [116].Chen Y, Durakoglugil MS, Xian X, Herz J, ApoE4 reduces glutamate receptor function and synaptic plasticity by selectively impairing ApoE receptor recycling, Proc Natl Acad Sci U S A 107 (2010) 12011–12016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [117].Leung L, Andrews-Zwilling Y, Yoon SY, Jain S, Ring K, Dai J, Wang MM, Tong L, Walker D, Huang Y, Apolipoprotein E4 causes age- and sex-dependent impairments of hilar GABAergic interneurons and learning and memory deficits in mice, PLoS One 7 (2012) e53569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [118].Tong LM, Yoon SY, Andrews-Zwilling Y, Yang A, Lin V, Lei H, Huang Y, Enhancing GABA Signaling during Middle Adulthood Prevents Age-Dependent GABAergic Interneuron Decline and Learning and Memory Deficits in ApoE4 Mice, J Neurosci 36 (2016) 2316–2322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [119].Nuriel T, Angulo SL, Khan U, Ashok A, Chen Q, Figueroa HY, Emrani S, Liu L, Herman M, Barrett G, Savage V, Buitrago L, Cepeda-Prado E, Fung C, Goldberg E, Gross SS, Hussaini SA, Moreno H, Small SA, Duff KE, Neuronal hyperactivity due to loss of inhibitory tone in APOE4 mice lacking Alzheimer’s disease-like pathology, Nature communications 8 (2017) 1464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [120].Gillespie AK, Jones EA, Lin YH, Karlsson MP, Kay K, Yoon SY, Tong LM, Nova P, Carr JS, Frank LM, Huang Y, Apolipoprotein E4 Causes Age-Dependent Disruption of Slow Gamma Oscillations during Hippocampal Sharp-Wave Ripples, Neuron 90 (2016) 740–751. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [121].Muenchhoff J, Song F, Poljak A, Crawford JD, Mather KA, Kochan NA, Yang Z, Trollor JN, Reppermund S, Maston K, Theobald A, Kirchner-Adelhardt S, Kwok JB, Richmond RL, McEvoy M, Attia J, Schofield PW, Brodaty H, Sachdev PS, Plasma apolipoproteins and physical and cognitive health in very old individuals, Neurobiol Aging 55 (2017) 49–60. [DOI] [PubMed] [Google Scholar]
  • [122].Nielsen HM, Chen K, Lee W, Chen Y, Bauer RJ 3rd, Reiman E, Caselli R, Bu G, Peripheral apoE isoform levels in cognitively normal APOE epsilon3/epsilon4 individuals are associated with regional gray matter volume and cerebral glucose metabolism, Alzheimer’s research & therapy 9 (2017) 5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [123].Lane-Donovan C, Wong WM, Durakoglugil MS, Wasser CR, Jiang S, Xian X, Herz J, Genetic Restoration of Plasma ApoE Improves Cognition and Partially Restores Synaptic Defects in ApoE-Deficient Mice, J Neurosci 36 (2016) 10141–10150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [124].Lane-Donovan C, Herz J, Is apolipoprotein e required for cognitive function in humans?: implications for Alzheimer drug development, JAMA neurology 71 (2014) 1213–1215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [125].Schaefer EJ, Gregg RE, Ghiselli G, Forte TM, Ordovas JM, Zech LA, Brewer HB Jr., Familial apolipoprotein E deficiency, J Clin Invest 78 (1986) 1206–1219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [126].Mak AC, Pullinger CR, Tang LF, Wong JS, Deo RC, Schwarz JM, Gugliucci A, Movsesyan I, Ishida BY, Chu C, Poon A, Kim P, Stock EO, Schaefer EJ, Asztalos BF, Castellano JM, Wyss-Coray T, Duncan JL, Miller BL, Kane JP, Kwok PY, Malloy MJ, Effects of the absence of apolipoprotein e on lipoproteins, neurocognitive function, and retinal function, JAMA neurology 71 (2014) 1228–1236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [127].Kim J, Eltorai AE, Jiang H, Liao F, Verghese PB, Kim J, Stewart FR, Basak JM, Holtzman DM, Anti-apoE immunotherapy inhibits amyloid accumulation in a transgenic mouse model of Abeta amyloidosis, J Exp Med 209 (2012) 2149–2156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [128].Liao F, Hori Y, Hudry E, Bauer AQ, Jiang H, Mahan TE, Lefton KB, Zhang TJ, Dearborn JT, Kim J, Culver JP, Betensky R, Wozniak DF, Hyman BT, Holtzman DM, Anti-ApoE antibody given after plaque onset decreases Abeta accumulation and improves brain function in a mouse model of Abeta amyloidosis, J Neurosci 34 (2014) 7281–7292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [129].Liao F, Li A, Xiong M, Bien-Ly N, Jiang H, Zhang Y, Finn MB, Hoyle R, Keyser J, Lefton KB, Robinson GO, Serrano JR, Silverman AP, Guo JL, Getz J, Henne K, Leyns CE, Gallardo G, Ulrich JD, Sullivan PM, Lerner EP, Hudry E, Sweeney ZK, Dennis MS, Hyman BT, Watts RJ, Holtzman DM, Targeting of nonlipidated, aggregated apoE with antibodies inhibits amyloid accumulation, J Clin Invest 128 (2018) 2144–2155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [130].Moutinho M, Landreth GE, Therapeutic potential of nuclear receptor agonists in Alzheimer’s disease, J Lipid Res 58 (2017) 1937–1949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [131].Tesseur I, Lo AC, Roberfroid A, Dietvorst S, Van Broeck B, Borgers M, Gijsen H, Moechars D, Mercken M, Kemp J, D’Hooge R, De Strooper B, Comment on “ApoE-directed therapeutics rapidly clear beta-amyloid and reverse deficits in AD mouse models”, Science 340 (2013) 924–e. [DOI] [PubMed] [Google Scholar]
  • [132].Veeraraghavalu K, Zhang C, Miller S, Hefendehl JK, Rajapaksha TW, Ulrich J, Jucker M, Holtzman DM, Tanzi RE, Vassar R, Sisodia SS, Comment on “ApoE-directed therapeutics rapidly clear beta-amyloid and reverse deficits in AD mouse models”, Science 340 (2013) 924–f. [DOI] [PubMed] [Google Scholar]
  • [133].Cummings JL, Zhong K, Kinney JW, Heaney C, Moll-Tudla J, Joshi A, Pontecorvo M, Devous M, Tang A, Bena J, Double-blind, placebo-controlled, proof-of-concept trial of bexarotene Xin moderate Alzheimer’s disease, Alzheimer’s research & therapy 8 (2016) 4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [134].Boehm-Cagan A, Michaelson DM, Reversal of apoE4-driven brain pathology and behavioral deficits by bexarotene, J Neurosci 34 (2014) 7293–7301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [135].Mounier A, Georgiev D, Nam KN, Fitz NF, Castranio EL, Wolfe CM, Cronican AA, Schug J, Lefterov I, Koldamova R, Bexarotene-Activated Retinoid X Receptors Regulate Neuronal Differentiation and Dendritic Complexity, J Neurosci 35 (2015) 11862–11876. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [136].Hong C, Tontonoz P, Liver X receptors in lipid metabolism: opportunities for drug discovery, Nat Rev Drug Discov 13 (2014) 433–444. [DOI] [PubMed] [Google Scholar]
  • [137].Tousi B, The emerging role of bexarotene in the treatment of Alzheimer’s disease: current evidence, Neuropsychiatric disease and treatment 11 (2015) 311–315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [138].Tai LM, Koster KP, Luo J, Lee SH, Wang YT, Collins NC, Ben Aissa M, Thatcher GR, LaDu MJ, Amyloid-beta pathology and APOE genotype modulate retinoid X receptor agonist activity in vivo, J Biol Chem 289 (2014) 30538–30555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [139].Handattu SP, Garber DW, Monroe CE, van Groen T, Kadish I, Nayyar G, Cao D, Palgunachari MN, Li L, Anantharamaiah GM, Oral apolipoprotein A-I mimetic peptide improves cognitive function and reduces amyloid burden in a mouse model of Alzheimer’s disease, Neurobiol Dis 34 (2009) 525–534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [140].Handattu SP, Monroe CE, Nayyar G, Palgunachari MN, Kadish I, van Groen T, Anantharamaiah GM, Garber DW, In vivo and in vitro effects of an apolipoprotein e mimetic peptide on amyloid-beta pathology, J Alzheimers Dis 36 (2013) 335–347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [141].Boehm-Cagan A, Bar R, Liraz O, Bielicki JK, Johansson JO, Michaelson DM, ABCA1 Agonist Reverses the ApoE4-Driven Cognitive and Brain Pathologies, J Alzheimers Dis 54 (2016) 1219–1233. [DOI] [PubMed] [Google Scholar]
  • [142].Laskowitz DT, Wang H, Chen T, Lubkin DT, Cantillana V, Tu TM, Kernagis D, Zhou G, Macy G, Kolls BJ, Dawson HN, Neuroprotective pentapeptide CN-105 is associated with reduced sterile inflammation and improved functional outcomes in a traumatic brain injury murine model, Scientific reports 7 (2017) 46461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [143].Vitek MP, Christensen DJ, Wilcock D, Davis J, Van Nostrand WE, Li FQ, Colton CA, APOE-mimetic peptides reduce behavioral deficits, plaques and tangles in Alzheimer’s disease transgenics, Neurodegener Dis 10 (2012) 122–126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [144].Ghosal K, Stathopoulos A, Thomas D, Phenis D, Vitek MP, Pimplikar SW, The apolipoprotein-E-mimetic COG112 protects amyloid precursor protein intracellular domain-overexpressing animals from Alzheimer’s disease-like pathological features, Neurodegener Dis 12 (2013) 51–58. [DOI] [PubMed] [Google Scholar]
  • [145].Montoliu-Gaya L, Mulder SD, Veerhuis R, Villegas S, Effects of an Abeta-antibody fragment on Abeta aggregation and astrocytic uptake are modulated by apolipoprotein E and J mimetic peptides, PLoS One 12 (2017) e0188191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [146].Montoliu-Gaya L, Mulder SD, Herrebout MAC, Baayen JC, Villegas S, Veerhuis R, Abeta-oligomer uptake and the resulting inflammatory response in adult human astrocytes are precluded by an anti-Abeta single chain variable fragment in combination with an apoE mimetic peptide, Mol Cell Neurosci 89 (2018) 49–59. [DOI] [PubMed] [Google Scholar]
  • [147].Bloedon LT, Dunbar R, Duffy D, Pinell-Salles P, Norris R, DeGroot BJ, Movva R, Navab M, Fogelman AM, Rader DJ, Safety, pharmacokinetics, and pharmacodynamics of oral apoA-I mimetic peptide D-4F in high-risk cardiovascular patients, J Lipid Res 49 (2008) 1344–1352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [148].Watson CE, Weissbach N, Kjems L, Ayalasomayajula S, Zhang Y, Chang I, Navab M, Hama S, Hough G, Reddy ST, Soffer D, Rader DJ, Fogelman AM, Schecter A, Treatment of patients with cardiovascular disease with L-4F, an apo-A1 mimetic, did not improve select biomarkers of HDL function, J Lipid Res 52 (2011) 361–373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [149].Dunbar RL, Movva R, Bloedon LT, Duffy D, Norris RB, Navab M, Fogelman AM, Rader DJ, Oral Apolipoprotein A-I Mimetic D-4F Lowers HDL-Inflammatory Index in High-Risk Patients: A First-in-Human Multiple-Dose, Randomized Controlled Trial, Clinical and translational science 10 (2017) 455–469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [150].Chernick D, Ortiz-Valle S, Jeong A, Swaminathan SK, Kandimalla KK, Rebeck GW, Li L, High-density lipoprotein mimetic peptide 4F mitigates amyloid-beta-induced inhibition of apolipoprotein E secretion and lipidation in primary astrocytes and microglia, J Neurochem 147 (2018) 647–662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [151].Chernick D, Ortiz-Valle S, Gram A, Li L, A clinically tested HDL mimetic peptide mitigates apoE4 associated lipidation and memory deficits, Program No. 746.11. 2018 Neuroscience Meeting Planner. San Diego, CA: Society for Neuroscience, 2018. Online. [Google Scholar]
  • [152].Mahley RW, Weisgraber KH, Huang Y, Apolipoprotein E4: a causative factor and therapeutic target in neuropathology, including Alzheimer’s disease, Proc Natl Acad Sci U S A 103 (2006) 5644–5651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [153].Zhong N, Weisgraber KH, Understanding the association of apolipoprotein E4 with Alzheimer disease: clues from its structure, J Biol Chem 284 (2009) 6027–6031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [154].Mahley RW, Huang Y, Small-molecule structure correctors target abnormal protein structure and function: structure corrector rescue of apolipoprotein E4-associated neuropathology, J Med Chem 55 (2012) 8997–9008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [155].Rosenberg JB, Kaplitt MG, De BP, Chen A, Flagiello T, Salami C, Pey E, Zhao L, Ricart Arbona RJ, Monette S, Dyke JP, Ballon DJ, Kaminsky SM, Sondhi D, Petsko GA, Paul SM, Crystal RG, AAVrh.10-Mediated APOE2 Central Nervous System Gene Therapy for APOE4-Associated Alzheimer’s Disease, Human gene therapy. Clinical development 29 (2018) 24–47. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [156].Agellon LB, Walsh A, Hayek T, Moulin P, Jiang XC, Shelanski SA, Breslow JL, Tall AR, Reduced high density lipoprotein cholesterol in human cholesteryl ester transfer protein transgenic mice, J Biol Chem 266 (1991) 10796–10801. [PubMed] [Google Scholar]
  • [157].Liao F, Zhang TJ, Jiang H, Lefton KB, Robinson GO, Vassar R, Sullivan PM, Holtzman DM, Murine versus human apolipoprotein E4: differential facilitation of and co-localization in cerebral amyloid angiopathy and amyloid plaques in APP transgenic mouse models, Acta neuropathologica communications 3 (2015) 70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [158].Maloney B, Ge YW, Alley GM, Lahiri DK, Important differences between human and mouse APOE gene promoters: limitation of mouse APOE model in studying Alzheimer’s disease, J Neurochem 103 (2007) 1237–1257. [DOI] [PubMed] [Google Scholar]
  • [159].Raffai RL, Dong LM, Farese RV Jr., Weisgraber KH, Introduction of human apolipoprotein E4 “domain interaction” into mouse apolipoprotein E, Proc Natl Acad Sci U S A 98 (2001) 11587–11591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [160].Tani M, Matera R, Horvath KV, Hasan TS, Schaefer EJ, Asztalos BF, The influence of apoE-deficiency and LDL-receptor-deficiency on the HDL subpopulation profile in mice and in humans, Atherosclerosis 233 (2014) 39–44. [DOI] [PubMed] [Google Scholar]

RESOURCES