NIH Public Access Author Manuscript Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

NIH-PA Author Manuscript

Published in final edited form as: Cogn Sci (Hauppauge). 2010 July 1; 5(1): 35–56.

CHOLESTEROL AND NEURONAL SUSCEPTIBILITY TO BETAAMYLOID TOXICITY Alexandra M. Nicholson and Adriana Ferreira* Department of Cell and Molecular Biology, Feinberg School of Medicine, Northwestern University, Chicago, IL 60611

Abstract

NIH-PA Author Manuscript

Alzheimer’s disease (AD) is a devastating neurocognitive disorder rapidly growing across the elderly population. Although few cases arise due to genetic mutations, sporadic AD is the most common form of this disease. Therefore, there is a continuing research effort to discover a unifying cause of this form of AD. To date, the only strong genetic correlate to the sporadic AD is inheritance of the apolipoprotein E4 (ApoE4) allele, whose encoded protein is involved in cholesterol transport in the central nervous system. This genetic link has prompted a series of studies on the potential molecular mechanisms by which cholesterol could modulate neuronal degeneration in the context of AD. In this review, we discussed the involvement of cholesterol in the production of the pathological hallmarks of the disease and how it might alter the susceptibility of cells to AD-related insult. Finally, we discussed the use of cholesterol-lowering drugs as a potential preventative approach in AD.

Keywords Cholesterol; Alzheimer’s disease; beta-amyloid; statins

Introduction NIH-PA Author Manuscript

Alzheimer’s disease (AD) is a progressive neurodegenerative disorder characterized by memory loss, language deficits, and space and time disorientation, among other neuropsychiatric symptoms (reviewed by Collie and Maruff, 2000). Since its discovery by Alois Alzheimer in 1901, AD has continually been of scientific interest due to its increasing effect among the elderly population. As a result, the morphological hallmarks of this neurodegenerative disorder, amyloid plaques and neurofibrillary tangles (NFTs), have been extensively studied. These lesions consist of extracellular deposits of the beta-amyloid (Aβ) peptide and intracellular accumulation of hyperphosphoylated forms of the microtubuleassociated protein tau, respectively (Glenner and Wong, 1984; Kosik et al., 1986; Wood et al., 1986). Though the multifaceted nature of AD has made it difficult to distinguish whether these pathological hallmarks initiate or are a consequence of the disease, understanding how

*

Correspondence should be addressed to: Adriana Ferreira, M.D., Ph. D., Cell and Molecular Biology Department Northwestern University, Ward Building 8-140, 303 East Chicago Avenue, Chicago, Illinois 60611, Phone: (312) 503-0597, Fax: (312) 503-7345, [email protected].

Nicholson and Ferreira

Page 2

NIH-PA Author Manuscript

these aggregates originate has been and remains a question on the forefront of AD research. A growing body of evidence suggests that the accumulation of Aβ peptides, formed by βand γ-secretase digestion of the amyloid precursor protein (APP), induces tau hyperphosphorylation and proteolysis leading to the formation of NFT (Vassar et al., 1999; Chung et al., 2001; Gamblin et al., 2003; reviewed by Tsai et al., 2004; Takashima, 2006). Unfortunately, researchers have been unsuccessful in developing pharmacological or therapeutic interventions that can halt the progression of protein dysfunction in AD. Therefore, it is becoming more critical to minimize the risk of developing the disease in an attempt to avoid its onset.

NIH-PA Author Manuscript

Cholesterol has gained interest as a potential risk factor for AD as a result of numerous studies. Early reports suggested a link between cardiovascular diseases and a risk of developing AD (Sparks et al., 1990; Kokmen et al., 1991; Sparks et al., 1993; Hofman et al., 1997). More recently, a population-based study demonstrated that high cholesterol levels represent a significant risk factor for AD (Kivipelto et al., 2001; Solomon et al., 2009). A relationship between moderately elevated cholesterol levels in midlife and increase incidence of AD and vascular dementia 3 decades later was also established in a large (9,844 participants) and diverse cohort (Solomon et al., 2009). Additionally, inheritance of the apolipoprotein E4 (ApoE4) allele, the transcript for a protein involved in cholesterol transport in the central nervous system (CNS), increases the probability one will develop AD (Corder et al., 1993). The picture emerging from these studies strongly suggests that cholesterol could play a role in the pathobiology of this disease. In this review, we discussed the potential role of cholesterol, including how it might be involved in the development of the pathological hallmarks of AD as well as how it might directly or indirectly intensify their mechanisms of neurotoxicity on a cellular level. Finally, we give an overview of the controversy behind cholesterol-lowering drugs and their potential to decrease one’s risk of developing AD.

Cholesterol production in the brain

NIH-PA Author Manuscript

Cholesterol constitutes a major component of the myelin sheath in the nervous system (reviewed by Dietschy, 2009). However, neurons of the hippocampus, the brain area primarily affected by AD pathology, are not myelinated (Andersen, 1960; Raastad and Shepherd, 2003). In non-myelinated cells, most of the cholesterol resides primarily in the plasma membrane where it is organized into specific membrane microdomains known as lipid rafts. These specific membrane compartments are involved in cellular processes, protein sorting, signaling complex formation, and the initiation of several signal transduction pathways (reviewed by Simons and Ikonen, 1997; Brown and London, 1998; Simons and Ikonen, 2000; Simons and Toomre, 2000; Ikonen, 2001; Suzuki, 2002). Cholesterol determines the proper cellular membrane fluidity, membrane protein components and their organization into lipid rafts, and is a critical precursor or cofactor for other signaling molecules. Therefore, the maintenance of proper cholesterol synthesis, transport, and intracellular sorting is tightly regulated. Most mammalian cells can synthesize cholesterol according to their own requirement, with acetate being the starting substrate. In cells outside of the CNS, most cholesterol originates either by de novo synthesis or from exogenously supplied lipoproteins secreted by the liver (Spady and Dietschy, 1983; Dietschy et al., 1993).

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 3

NIH-PA Author Manuscript NIH-PA Author Manuscript

Since brain cholesterol levels are five to ten times higher than those in other organs and cholesterol turnover in the brain is 6 times slower than in other tissues, it is not surprising that cholesterol homeostasis in brain cells is distinct from other cell types (Bjorkhem et al., 1997; reviewed by Dietschy and Turley, 2004). To date, there is solid evidence that the majority of cholesterol in the CNS is synthesized locally rather than imported from blood lipoproteins (Dietschy et al., 1983; Edmond et al., 1991; Jurevics and Morell, 1995; Osono et al., 1995; Turley et al., 1998). Central neurons express 3-hydroxy-3-methyl-glutaryl-CoA (HMG-CoA) reductase, the enzyme required in the cholesterol synthesis pathway (Volpe and Hennessy, 1977; Swanson et al., 1988). Furthermore, cholesterol synthesis has been detected in both embryonic and postnatal neuron cultures (Lopes-Cardozo et al., 1986; Saito et al., 1987; Tabernero et al., 1993). However, it has been questioned whether the amount of HMG-CoA reductase is enough to establish the necessary levels of cholesterol, as well as other HMG-CoA reductase-dependent molecules, as neurons develop. This question is in agreement with multiple reports suggesting that neurons require glial-derived cholesterol for growth and survival (Ignatius et al., 1986; Handelmann et al., 1992; Poirier et al., 1993; Meyer-Franke et al., 1995; Mauch et al., 2001; reviewed by Goritz et al., 2002; Pfrieger, 2003b, 2003a). Glial cholesterol, obtained either by de novo synthesis or by recycling cholesterol that has been released from degenerating nerve terminals, complexes with lipidcarrying molecules to produce cholesterol-containing low density lipoprotein (LDL) particles which are secreted by glial cells (Shanmugaratnam et al., 1997). One of the most abundant of these lipid carriers is the glia-derived ApoE (Elshourbagy et al., 1985). ApoE proteins are involved in the transport and distribution of cholesterol and other lipids, particularly to and from neuronal subcellular compartments (reviewed by Mahley, 1988; Poirier, 1994; Herz and Beffert, 2000). On the surface of neuronal cells, ApoE-containing lipoproteins are ligands primarily for the LDL family of cell surface receptors, which are then internalized and broken down within the cell to free the cholesterol from its associated lipoprotein complex (Rebeck et al., 1993; Bu et al., 1994).

ApoE4 in cholesterol homeostasis and AD pathogenesis

NIH-PA Author Manuscript

Sporadic AD, unlike the familial form, is the most common form of the disease and lacks genetic mutations as a primary cause. However, there is one gene mapped to chromosome 19 that has been implicated as a significant risk factor for both forms of AD (Pericak-Vance et al., 1991). This gene encodes the ApoE protein mentioned above as critical for CNS cholesterol homeostasis. This lipid carrier is polymorphic with three alleles encoding different isoforms: ApoE2, ApoE3, and ApoE4, which differ only in their amino acid residues at positions 112 and 158 (reviewed by Mahley, 1988). Though ApoE3 is most abundantly expressed, it remains inconclusive whether this isoform might be involved in AD pathogenesis. Interestingly, ApoE2, whose allelic frequency is the lowest of the three isoforms in the United States population, seems to have potential neuroprotective qualities that actually may reduce the risk of acquiring AD (Corder et al., 1994; Lippa et al., 1997; Lee et al., 2002). Conversely, the ApoE4 polymorphism was identified as an important factor to the etiology of many AD cases (Corder et al., 1993). This study examined the ApoE genotype profile of individuals from over 40 families and assessed the proportion of AD-affected individuals with each genotype. Their results established that only 20% of

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 4

NIH-PA Author Manuscript

individuals containing either the ApoE2/3 or ApoE3/3 genotype had AD. Forty-seven percent of people with one ApoE4 allele were affected with AD, while approximately 90% of individuals carrying two ApoE4 alleles had been pathologically diagnosed with this disease. Not only did the number of inherited ApoE4 alleles correlate with an increased risk of developing the disease, but also with a decrease in the average age of onset (Corder et al., 1993).

NIH-PA Author Manuscript

The data reviewed above suggested that ApoE isoforms might differentially contribute to the pathobiology of AD on a cellular level, though the mechanism of this correlation is not completely understood. Studies have shown that the ApoE4 isoform has a greater binding affinity for LDL receptors than the other ApoE isoforms (Dong et al., 1994; Dong and Weisgraber, 1996). Furthermore, ApoE4 was the least potent isoform to promote cholesterol efflux from both cortical neurons and glial cells (Michikawa et al., 2000). As a result of these findings, research efforts assessed to what extent ApoE4 might promote an increase in neuron cholesterol levels, rendering them more susceptible to either the production or the toxicity induced by the proteinaceous deposits typical in AD. Evidence has emerged suggesting that ApoE does, indeed, regulate cholesterol levels in the synaptic plasma membranes (SPM) (Igbavboa et al., 1997). In a study conducted in 2002, Hayashi and colleagues utilized ApoE4 knock-in mice to determine cholesterol content in the SPMs isolated from these transgenic mice, as well as in ApoE3-overexpressing and wild type mice. Unlike ApoE3 and wild type mice, animals overexpressing ApoE4 had increased cholesterol in the exofacial leaflet of SPMs at 8 weeks of age (Hayashi et al., 2002). Although these results provided evidence that ApoE4 might be linked to AD in a cholesterol-dependent manner, the effects of ApoE overexpression are vast and will require more analyses to elucidate which ApoE4-induced cellular modifications primarily contribute to AD onset and pathology.

Cholesterol and Aβ production

NIH-PA Author Manuscript

One of the early steps in the pathobiology of AD is the increase in Aβ levels (reviewed by Bharadwaj et al., 2009). The production of this toxic peptide is the result of the amyloidogenic secretase cleavage of the APP transmembrane protein. To initiate Aβ formation, APP is first cleaved by β-secretase, an enzyme more commonly known as β-site APP cleaving enzyme or BACE-1 (Urmoneit et al., 1995; Vassar et al., 1999). As a result of this cleavage event, the extracellular portion of APP is released in a soluble form, APPsβ. The remaining portion of the APP molecule within the membrane is further cleaved at the Cterminal end of Aβ by γ-secretase to yield the Aβ protein primarily of 40–42 amino acids (Anderson et al., 1992; Haass et al., 1992; Roher et al., 1993; Xia et al., 1997). Alternatively, APP undergoing non-amyloidogenic processing is cleaved first by the enzyme α-secretase followed by γ-secretase. The α-secretase cleavage event initiates the release a different soluble portion of APP than that of β-secretase APP cleavage, APPsα. Furthermore, APP cleavage by α-secretase is considered non-amyloidogenic APP processing because the α-secretase cleavage site is within the Aβ portion of APP (Sisodia et al., 1990). Thus, factors that favor either α- or β-secretase APP proteolysis could decrease or increase the Aβ burden, respectively.

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 5

NIH-PA Author Manuscript

Cholesterol seems to be one of those factors. Earlier reports showed that cholesterol-linked cardiovascular diseases, such as coronary artery disease, were associated with Aβ deposition in the brain similar to that of AD patients (Sparks et al., 1990; Sparks et al., 1993). However, it was not until 1994 when the first experimental evidence suggesting interplay between cholesterol consumption and Aβ formation in the brain was published. In this study, rabbits were fed either a control or a hypercholesterol diet for up to 8 weeks before they were sacrificed and their brains analyzed. Rabbits on the hypercholesterol diet had significantly elevated plasma cholesterol levels as well as clinical signs of jaundice and renal failure; most likely do to fatty liver and fatty deposits in their arteries. Interestingly, a timedependent increase in Aβ immunoreactivity was detected in the hippocampus and cortex of rabbits on the hypercholesterol diet (Sparks et al., 1994). Similar observations have since been obtained in transgenic mouse models of AD (Howland et al., 1998; Refolo et al., 2000; Shie et al., 2002).

NIH-PA Author Manuscript

Although a debate remains as to whether plasma cholesterol might affect levels of this sterol in the brain, these findings provided significant evidence that cholesterol might be linked to Aβ production and deposition. Further support for such a role was obtained by manipulating membrane cholesterol in different cell types to determine to what extent Aβ production, or that of similar proteins, is altered. For example, increasing membrane cholesterol by means of incubations with cholesterol solubilized in either a methyl-β-cyclodextrin (MBCD) complex or by ethanol in APP-transfected HEK293 cells decreased the amount of APPsα secreted (Bodovitz and Klein, 1996). Similar results were obtained using a different approach to increase cellular cholesterol content in COS-1 cells. In these experiments, COS-1 cells were exposed to various forms of LDL containing non-esterified cholesterol. While the dose of cholesterol-containing LDL complexes correlated with an increase in membrane cholesterol content in these cells, a subsequent negative correlation in APPsα release was observed (Racchi et al., 1997).

NIH-PA Author Manuscript

Both of these studies suggested that increased membrane cholesterol may attenuate αsecretase processing of APP, though neither provided evidence of the converse: that decreasing membrane cholesterol increases α-secretase activity. This information arose when research efforts assessed α-secretase activity in response to decreased membrane cholesterol in both neuronal and non-neuronal cell lines. In these experiments, membrane cholesterol was removed from cellular membranes using MBCD, a glycoprotein able to encapsulate cholesterol to render it water-soluble. Cells in which cholesterol had been reduced by MBCD showed a marked increase in secreted APPsα (Kojro et al., 2001). Additionally, decreased cholesterol actually reduced the amount of Aβ peptides generated in these cellular models. These data supported previous findings in cultured hippocampal neurons (Simons et al., 1998). Taken together, these results provided insights into the role of cellular cholesterol levels and the regulation of the enzymes involved in APP processing leading to Aβ production in the context of AD.

Cholesterol might mediate Aβ-membrane interactions Not only is the process of Aβ production a toxic factor in AD pathogenesis, but also the effect that Aβ has on neuronal cells after it has been released into the extracellular space.

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 6

NIH-PA Author Manuscript NIH-PA Author Manuscript

Although different hypotheses remain as to how exogenous Aβ exerts its neurotoxic effects, most suggest the interaction of Aβ with surrounding cellular membranes. In fact, using electron microscopy techniques, Aβ has been shown to associate with cell surface plasma membranes in the brains of AD patients (Yamaguchi et al., 2000). Similar findings were reported in aged dogs, since dogs are a species of vertebrates that demonstrate agedependent cognitive decline correlated with Aβ accumulation (Torp et al., 2000). Aβ has also been shown to exert membrane-disordering properties, both in cellular and artificial membranes, further supporting an interaction between this peptide and membrane structures. Several reports have provided evidence of Aβ binding to cholesterol-rich low-density areas of membranes (Morishima-Kawashima and Ihara, 1998; Mizuno et al., 1999; Oshima et al., 2001; Kawarabayashi et al., 2004). One of the first biochemical assessments of Aβ-lipid interaction was conducted by Avdulov et al. (1997). This research group discovered a preferential interaction of Aβ with cholesterol over other lipids of the membrane bilayer using in vitro techniques (Avdulov et al., 1997). Later evidence was obtained of Aβ binding to other components of the membrane, such as ganglioside-M1 (GM1). Interestingly, the amount of GM1 necessary for Aβ binding was decreased when cholesterol was present, an environment mimicking that of actual lipid rafts (Matsuzaki and Horikiri, 1999). Data challenging the enhanced effects of cholesterol on Aβ-membrane interactions have also been reported. For example, membrane cholesterol in synthetic membranes actually excluded Aβ peptide from the membrane, most likely due to rigidifying nature of membrane cholesterol (Curtain et al., 2003). These results were in agreement with a study done in 2006 in which the addition of cholesterol to synthetic membranes reduced the insertion of Aβ25–35 (Dante et al., 2006). These conflicting results on the role of cholesterol in Aβ-membrane interactions could be due to the model system in which these questions were addressed. Regardless, it is difficult to dispute that cholesterol might, at least in part, mediate the ability of Aβ to interact with its surrounding cellular membranes.

Cholesterol and Aβ toxicity

NIH-PA Author Manuscript

The in vitro models of cellular membranes used in the experiments described above provided useful data regarding the role of cholesterol as a modulator of Aβ-membrane interactions. However, they did not address to what extent cholesterol might alter the susceptibility of cells to Aβ toxicity at a cellular level in the context of AD. It is known that Aβ exerts its neurotoxic effects, at least in part, by enhancing calcium (Ca2+) influx (reviewed by Ingram, 2005; Smith et al., 2005; Mattson, 2007; Bojarski et al., 2008; Green and LaFerla, 2008). Interestingly, cholesterol has been implicated in the susceptibility of cells to Ca2+ influx (Bastiaanse et al., 1994). Cholesterol appeared to be protective against Aβ-induced Ca2+ influx both in dissociated mouse brain cells and in human lymphocytes and immortalized hypothalamic neurons (Hartmann et al., 1994; Kawahara and Kuroda, 2001). This seems not to be the case in mature hippocampal cultures. In this report, elevated membrane cholesterol actually increased the susceptibility of cells to Aβ-induced elevation in Ca2+ influx leading to cell death via a tau-dependent mechanism (Nicholson and Ferreira, 2009). Thus, the effects of cholesterol on Ca2+ influx induced by Aβ might vary in different cell types and, more importantly, in different stages of neuronal maturation.

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 7

NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Another mechanism through which Aβ is thought to cause neurotoxicity is by means of inducing abnormal posttranslational processing of the tau protein. Of these posttranslational modifications, Aβ-induced tau hyperphosphorylation has been most extensively studied due to its hypothesized role in tau aggregation and NFT formation in AD (reviewed by Lee, 1996; Michaelis et al., 2002; Chun and Johnson, 2007; Iqbal et al., 2009). However, recent research efforts have also shown that Aβ-induced cleavage of tau can not only accelerate tau aggregation, but can generate tau fragments that are neurotoxic themselves (Chung et al., 2001; Gamblin et al., 2003; Park and Ferreira, 2005; Park et al., 2007). The suspicion that cholesterol might be a factor in AD tau pathology arose in studies conducted in models of Niemann-Pick type C (NPC) disease. Patients with this disease show disruption in several cellular transport processes leading to the accumulation of cholesterol and glycolipids in lysosomal compartments. Interestingly, tau tangle pathology is found in the brains of NPC patients, with or without other AD histopathological changes (Auer et al., 1995; Love et al., 1995; Suzuki et al., 1995). One of the first reports suggesting a link between cholesterol and tau pathology on a cellular level determined that cellular cholesterol deficiency in primary cortical neurons induced tau phosphorylation in culture (Fan et al., 2001). Unfortunately, little is known regarding how cholesterol might be a factor in these cellular processes in the specific context of AD. Transgenic models of AD have been a useful tool in the beginning stages of understanding the relationship of many factors involved in this disease. Thus, phosphorylated tau has been detected to accumulate in cholesterol-rich portions of cellular membranes in mice overexpressing APP (Kawarabayashi et al., 2004). However, it remains unknown whether cholesterol was a direct mediator of these observations. In addition, when neuron cholesterol content was increased in transgenic mice fed hypercholesterol diets, enhanced tau kinase activation and subsequent tau phosphorylation were detected as compared to controls (Ghribi et al., 2006). A positive correlation between cholesterol content and tau toxicity was also established in primary hippocampal neurons in culture. When membrane cholesterol content in these cells was increased, either by the natural effect of aging or pharmacologically, their susceptibility to Aβ-induced calpain-mediated cleavage of tau into a neurotoxic fragment was significantly enhanced (Nicholson and Ferreira, 2009). Furthermore, the reciprocal was true in which decreasing membrane cholesterol content in hippocampal neurons attenuated tau cleavage induced by Aβ (Nicholson and Ferreira, 2009). Taken together, these results suggest that cellular cholesterol levels might be involved in the molecular mechanisms that underlie Aβ-induced excytotoxicity and tau posttranslational modifications.

Cholesterol-lowering drugs: a possibility for AD prevention? The data reviewed above strongly suggest a potential involvement of cholesterol in AD pathology. These findings raise the question as to whether cholesterol-lowering drugs might be beneficial for AD prevention and/or treatment. Although no drugs currently exist to specifically target cholesterol in the brain, statins have been commonly used to decrease peripheral cholesterol levels. Statins are a class of drugs that inhibit HMG-CoA reductase, prevent the formation and entry of LDL cholesterol into the circulation, and upregulate LDL receptor activity (Mabuchi et al., 1981; Bilheimer et al., 1983; Mabuchi et al., 1983). To what extent changes in peripheral cholesterol levels affect brain cholesterol is an open

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 8

NIH-PA Author Manuscript NIH-PA Author Manuscript

question. Regardless, longitudinal and case studies have been conducted to assess whether HMG-CoA reductase inhibitors might affect one’s risk of AD and its hallmark pathology (Jick et al., 2000; Wolozin et al., 2000; Scott and Laake, 2001; Tokuda et al., 2001; Buxbaum et al., 2002; Hajjar et al., 2002; Rockwood et al., 2002). Some AD clinical trials suggest that the use of statins could reduce the incidence of AD. Thus, preliminary results from the AD cholesterol lowering treatment trial showed the benefits of atorvastatin in cognitive function, clinical efficacy, and psychiatric symptoms after taking the drug for 12 months (Sparks et al., 2005). A decrease in the overall risk of developing AD and related dementias was also observed when individuals of 50 years and older were prescribed simvastatin, pravastatin, fluvastatin, lovostatin, or cerivastain (Jick et al., 2000; Wolozin et al., 2000; Buxbaum et al., 2002; Hajjar et al., 2002). Other clinical reports question a link between the use of these cholesterol-lowering drugs and AD (Scott and Laake, 200; Rockwood et al., 2002). These conflicting clinical outcomes might be due to the plieotropic nature of HMG-CoA reductase inhibitors, due to the requirement of this enzyme in the synthesis of other cellular molecules. Nor can it can be ruled out that the beneficial effects of statins could be independent of potential changes in brain cholesterol levels since these drugs also have anti-inflamatory, antioxidative, antithrombogenic, and immunological effects. In sum, even after extensive epidemiological, cellular, and biochemical investigations into the involvement of cholesterol in AD pathophysiology, its exact role in this disease remains obscure. Nevertheless, further investigation into the correlation between cholesterol and AD pathology could identify new targets for therapeutic intervention for the prevention and or treatment of this devastating disease.

Acknowledgments Supported by NIH grant RO1 NS39080 and Alzheimer’s Association grant 57869 to A.F.

REFERENCES

NIH-PA Author Manuscript

Andersen P. Interhippocampal impulses. II. Apical dendritic activation of CAI neurons. Acta Physiol Scand. 1960; 48:178–208. [PubMed: 13793336] Anderson JP, Chen Y, Kim KS, Robakis NK. An alternative secretase cleavage produces soluble Alzheimer amyloid precursor protein containing a potentially amyloidogenic sequence. J Neurochem. 1992; 59:2328–2331. [PubMed: 1431910] Auer IA, Schmidt ML, Lee VM, Curry B, Suzuki K, Shin RW, Pentchev PG, Carstea ED, Trojanowski JQ. Paired helical filament tau (PHFtau) in Niemann-Pick type C disease is similar to PHFtau in Alzheimer's disease. Acta Neuropathol. 1995; 90:547–551. [PubMed: 8615074] Avdulov NA, Chochina SV, Igbavboa U, Warden CS, Vassiliev AV, Wood WG. Lipid binding to amyloid beta-peptide aggregates: preferential binding of cholesterol as compared with phosphatidylcholine and fatty acids. Journal of Neurochemistry. 1997; 69:1746–1752. [PubMed: 9326304] Bastiaanse EM, Atsma DE, Kuijpers MM, Van der Laarse A. The effect of sarcolemmal cholesterol content on intracellular calcium ion concentration in cultured cardiomyocytes. Arch Biochem Biophys. 1994; 313:58–63. [PubMed: 8053687] Bharadwaj PR, Dubey AK, Masters CL, Martins RN, Macreadie IG. Abeta aggregation and possible implications in Alzheimer's disease pathogenesis. J Cell Mol Med. 2009; 13:412–421. [PubMed: 19374683]

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 9

NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Bilheimer DW, Grundy SM, Brown MS, Goldstein JL. Mevinolin and colestipol stimulate receptormediated clearance of low density lipoprotein from plasma in familial hypercholesterolemia heterozygotes. Proc Natl Acad Sci U S A. 1983; 80:4124–4128. [PubMed: 6575399] Bjorkhem I, Lutjohann D, Breuer O, Sakinis A, Wennmalm A. Importance of a novel oxidative mechanism for elimination of brain cholesterol. Turnover of cholesterol and 24(S)hydroxycholesterol in rat brain as measured with 18O2 techniques in vivo and in vitro. J Biol Chem. 1997; 272:30178–30184. [PubMed: 9374499] Bodovitz S, Klein WL. Cholesterol modulates alpha-secretase cleavage of amyloid precursor protein. J Biol Chem. 1996; 271:4436–4440. [PubMed: 8626795] Bojarski L, Herms J, Kuznicki J. Calcium dysregulation in Alzheimer's disease. Neurochem Int. 2008; 52:621–633. [PubMed: 18035450] Brown DA, London E. Functions of lipid rafts in biological membranes. Annu Rev Cell Dev Biol. 1998; 14:111–136. [PubMed: 9891780] Bu G, Maksymovitch EA, Nerbonne JM, Schwartz AL. Expression and function of the low density lipoprotein receptor-related protein (LRP) in mammalian central neurons. J Biol Chem. 1994; 269:18521–18528. [PubMed: 7518435] Buxbaum JD, Cullen EI, Friedhoff LT. Pharmacological concentrations of the HMG-CoA reductase inhibitor lovastatin decrease the formation of the Alzheimer beta-amyloid peptide in vitro and in patients. Front Biosci. 2002; 7:a50–a59. [PubMed: 11900994] Chun W, Johnson GV. The role of tau phosphorylation and cleavage in neuronal cell death. Front Biosci. 2007; 12:733–756. [PubMed: 17127334] Chung CW, Song YH, Kim IK, Yoon WJ, Ryu BR, Jo DG, Woo HN, Kwon YK, Kim HH, Gwag BJ, Mook-Jung IH, Jung YK. Proapoptotic effects of tau cleavage product generated by caspase-3. Neurobiol Dis. 2001; 8:162–172. [PubMed: 11162250] Collie A, Maruff P. The neuropsychology of preclinical Alzheimer's disease and mild cognitive impairment. Neurosci Biobehav Rev. 2000; 24:365–374. [PubMed: 10781696] 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. 1993; 261:921–923. [PubMed: 8346443] Corder EH, Saunders AM, Risch NJ, Strittmatter WJ, Schmechel DE, Gaskell PC Jr, Rimmler JB, Locke PA, Conneally PM, Schmader KE, et al. Protective effect of apolipoprotein E type 2 allele for late onset Alzheimer disease. Nat Genet. 1994; 7:180–184. [PubMed: 7920638] Curtain CC, Ali FE, Smith DG, Bush AI, Masters CL, Barnham KJ. Metal ions, pH, and cholesterol regulate the interactions of Alzheimer's disease amyloid-beta peptide with membrane lipid. J Biol Chem. 2003; 278:2977–2982. [PubMed: 12435742] Dante S, Hauss T, Dencher NA. Cholesterol inhibits the insertion of the Alzheimer's peptide Abeta(25–35) in lipid bilayers. Eur Biophys J. 2006; 35:523–531. [PubMed: 16670880] Dietschy JM. Central nervous system: cholesterol turnover, brain development and neurodegeneration. Biol Chem. 2009; 390:287–293. [PubMed: 19166320] 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. 2004; 45:1375– 1397. [PubMed: 15254070] Dietschy JM, Turley SD, Spady DK. Role of liver in the maintenance of cholesterol and low density lipoprotein homeostasis in different animal species, including humans. J Lipid Res. 1993; 34:1637–1659. [PubMed: 8245716] Dietschy JM, Kita T, Suckling KE, Goldstein JL, Brown MS. Cholesterol synthesis in vivo and in vitro in the WHHL rabbit, an animal with defective low density lipoprotein receptors. J Lipid Res. 1983; 24:469–480. [PubMed: 6304219] Dong LM, Weisgraber KH. 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. [PubMed: 8702576] Dong LM, Wilson C, Wardell MR, Simmons T, Mahley RW, Weisgraber KH, Agard DA. 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. [PubMed: 8071364]

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 10

NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Edmond J, Korsak RA, Morrow JW, Torok-Both G, Catlin DH. Dietary cholesterol and the origin of cholesterol in the brain of developing rats. J Nutr. 1991; 121:1323–1330. [PubMed: 1880610] Elshourbagy NA, Liao WS, Mahley RW, Taylor JM. Apolipoprotein E mRNA is abundant in the brain and adrenals, as well as in the liver, and is present in other peripheral tissues of rats and marmosets. Proc Natl Acad Sci U S A. 1985; 82:203–207. [PubMed: 3918303] Fan QW, Yu W, Senda T, Yanagisawa K, Michikawa M. Cholesterol-dependent modulation of tau phosphorylation in cultured neurons. J Neurochem. 2001; 76:391–400. [PubMed: 11208902] Gamblin TC, Chen F, Zambrano A, Abraha A, Lagalwar S, Guillozet AL, Lu M, Fu Y, Garcia-Sierra F, LaPointe N, Miller R, Berry RW, Binder LI, Cryns VL. Caspase cleavage of tau: linking amyloid and neurofibrillary tangles in Alzheimer's disease. Proc Natl Acad Sci U S A. 2003; 100:10032–10037. [PubMed: 12888622] Ghribi O, Larsen B, Schrag M, Herman MM. High cholesterol content in neurons increases BACE, beta-amyloid, and phosphorylated tau levels in rabbit hippocampus. Exp Neurol. 2006; 200:460– 467. [PubMed: 16696972] Glenner GG, Wong CW. Alzheimer's disease: initial report of the purification and characterization of a novel cerebrovascular amyloid protein. Biochem Biophys Res Commun. 1984; 120:885–890. [PubMed: 6375662] Goritz C, Mauch DH, Nagler K, Pfrieger FW. Role of glia-derived cholesterol in synaptogenesis: new revelations in the synapse-glia affair. J Physiol Paris. 2002; 96:257–263. [PubMed: 12445904] Green KN, LaFerla FM. Linking calcium to Abeta and Alzheimer's disease. Neuron. 2008; 59:190– 194. [PubMed: 18667147] Haass C, Schlossmacher MG, Hung AY, Vigo-Pelfrey C, Mellon A, Ostaszewski BL, Lieberburg I, Koo EH, Schenk D, Teplow DB, et al. Amyloid beta-peptide is produced by cultured cells during normal metabolism. Nature. 1992; 359:322–325. [PubMed: 1383826] Hajjar I, Schumpert J, Hirth V, Wieland D, Eleazer GP. The impact of the use of statins on the prevalence of dementia and the progression of cognitive impairment. J Gerontol A Biol Sci Med Sci. 2002; 57:M414–M418. [PubMed: 12084801] Handelmann GE, Boyles JK, Weisgraber KH, Mahley RW, Pitas RE. Effects of apolipoprotein E, beta-very low density lipoproteins, and cholesterol on the extension of neurites by rabbit dorsal root ganglion neurons in vitro. J Lipid Res. 1992; 33:1677–1688. [PubMed: 1464751] Hartmann H, Eckert A, Muller WE. Apolipoprotein E and cholesterol affect neuronal calcium signalling: the possible relationship to beta-amyloid neurotoxicity. Biochem Biophys Res Commun. 1994; 200:1185–1192. [PubMed: 8185566] Hayashi H, Igbavboa U, Hamanaka H, Kobayashi M, Fujita SC, Wood WG, Yanagisawa K. Cholesterol is increased in the exofacial leaflet of synaptic plasma membranes of human apolipoprotein E4 knock-in mice. Neuroreport. 2002; 13:383–386. [PubMed: 11930145] Herz J, Beffert U. Apolipoprotein E receptors: linking brain development and Alzheimer's disease. Nat Rev Neurosci. 2000; 1:51–58. [PubMed: 11252768] Hofman A, Ott A, Breteler MM, Bots ML, Slooter AJ, van Harskamp F, van Duijn CN, Van Broeckhoven C, Grobbee DE. Atherosclerosis, apolipoprotein E, and prevalence of dementia and Alzheimer's disease in the Rotterdam Study. Lancet. 1997; 349:151–154. [PubMed: 9111537] Howland DS, Trusko SP, Savage MJ, Reaume AG, Lang DM, Hirsch JD, Maeda N, Siman R, Greenberg BD, Scott RW, Flood DG. Modulation of secreted beta-amyloid precursor protein and amyloid beta-peptide in brain by cholesterol. J Biol Chem. 1998; 273:16576–16582. [PubMed: 9632729] Igbavboa U, Avdulov NA, Chochina SV, Wood WG. Transbilayer distribution of cholesterol is modified in brain synaptic plasma membranes of knockout mice deficient in the low-density lipoprotein receptor, apolipoprotein E, or both proteins. J Neurochem. 1997; 69:1661–1667. [PubMed: 9326295] Ignatius MJ, Gebicke-Harter PJ, Skene JH, Schilling JW, Weisgraber KH, Mahley RW, Shooter EM. Expression of apolipoprotein E during nerve degeneration and regeneration. Proc Natl Acad Sci U S A. 1986; 83:1125–1129. [PubMed: 2419900] Ikonen E. Roles of lipid rafts in membrane transport. Curr Opin Cell Biol. 2001; 13:470–477. [PubMed: 11454454]

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 11

NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Ingram VM. The role of Alzheimer Abeta peptides in ion transport across cell membranes. Subcell Biochem. 2005; 38:339–349. [PubMed: 15709487] Iqbal K, Liu F, Gong CX, Alonso Adel C, Grundke-Iqbal I. Mechanisms of tau-induced neurodegeneration. Acta Neuropathol. 2009; 118:53–69. [PubMed: 19184068] Jick H, Zornberg GL, Jick SS, Seshadri S, Drachman DA. Statins and the risk of dementia. Lancet. 2000; 356:1627–1631. [PubMed: 11089820] Jurevics H, Morell P. Cholesterol for synthesis of myelin is made locally, not imported into brain. J Neurochem. 1995; 64:895–901. [PubMed: 7830084] Kawahara M, Kuroda Y. Intracellular calcium changes in neuronal cells induced by Alzheimer's betaamyloid protein are blocked by estradiol and cholesterol. Cell Mol Neurobiol. 2001; 21:1–13. [PubMed: 11440193] Kawarabayashi T, Shoji M, Younkin LH, Wen-Lang L, Dickson DW, Murakami T, Matsubara E, Abe K, Ashe KH, Younkin SG. Dimeric amyloid beta protein rapidly accumulates in lipid rafts followed by apolipoprotein E and phosphorylated tau accumulation in the Tg2576 mouse model of Alzheimer's disease. J Neurosci. 2004; 24:3801–3809. [PubMed: 15084661] Kivipelto M, Helkala EL, Hanninen T, Laakso MP, Hallikainen M, Alhainen K, Soininen H, Tuomilehto J, Nissinen A. Midlife vascular risk factors and late-life mild cognitive impairment: A population-based study. Neurology. 2001; 56:1683–1689. [PubMed: 11425934] Kojro E, Gimpl G, Lammich S, Marz W, Fahrenholz F. Low cholesterol stimulates the nonamyloidogenic pathway by its effect on the alpha -secretase ADAM 10. Proc Natl Acad Sci U S A. 2001; 98:5815–5820. [PubMed: 11309494] Kokmen E, Beard CM, Chandra V, Offord KP, Schoenberg BS, Ballard DJ. Clinical risk factors for Alzheimer's disease: a population-based case-control study. Neurology. 1991; 41:1393–1397. [PubMed: 1891088] Kosik KS, Joachim CL, Selkoe DJ. Microtubule-associated protein tau (tau) is a major antigenic component of paired helical filaments in Alzheimer disease. Proc Natl Acad Sci U S A. 1986; 83:4044–4048. [PubMed: 2424016] Lee G, Pollard HB, Arispe N. Annexin 5 and apolipoprotein E2 protect against Alzheimer's amyloidbeta-peptide cytotoxicity by competitive inhibition at a common phosphatidylserine interaction site. Peptides. 2002; 23:1249–1263. [PubMed: 12128082] Lee VM. Regulation of tau phosphorylation in Alzheimer's disease. Ann N Y Acad Sci. 1996; 777:107–113. [PubMed: 8624072] Lippa CF, Smith TW, Saunders AM, Hulette C, Pulaski-Salo D, Roses AD. Apolipoprotein E-epsilon 2 and Alzheimer's disease: genotype influences pathologic phenotype. Neurology. 1997; 48:515– 519. [PubMed: 9040748] Lopes-Cardozo M, Larsson OM, Schousboe A. Acetoacetate and glucose as lipid precursors and energy substrates in primary cultures of astrocytes and neurons from mouse cerebral cortex. J Neurochem. 1986; 46:773–778. [PubMed: 3081684] Love S, Bridges LR, Case CP. Neurofibrillary tangles in Niemann-Pick disease type C. Brain. 1995; 118(Pt 1):119–129. [PubMed: 7894998] Mabuchi H, Sakai T, Sakai Y, Yoshimura A, Watanabe A, Wakasugi T, Koizumi J, Takeda R. Reduction of serum cholesterol in heterozygous patients with familial hypercholesterolemia. Additive effects of compactin and cholestyramine. N Engl J Med. 1983; 308:609–613. [PubMed: 6828091] Mabuchi H, Haba T, Tatami R, Miyamoto S, Sakai Y, Wakasugi T, Watanabe A, Koizumi J, Takeda R. Effect of an inhibitor of 3-hydroxy-3-methyglutaryl coenzyme A reductase on serum lipoproteins and ubiquinone-10-levels in patients with familial hypercholesterolemia. N Engl J Med. 1981; 305:478–482. [PubMed: 7254297] Mahley RW. Apolipoprotein E: cholesterol transport protein with expanding role in cell biology. Science. 1988; 240:622–630. [PubMed: 3283935] Matsuzaki K, Horikiri C. Interactions of amyloid beta-peptide (1–40) with ganglioside-containing membranes. Biochemistry. 1999; 38:4137–4142. [PubMed: 10194329] Mattson MP. Calcium and neurodegeneration. Aging Cell. 2007; 6:337–350. [PubMed: 17328689]

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 12

NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Mauch DH, Nagler K, Schumacher S, Goritz C, Muller EC, Otto A, Pfrieger FW. CNS synaptogenesis promoted by glia-derived cholesterol. Science. 2001; 294:1354–1357. [PubMed: 11701931] Meyer-Franke A, Kaplan MR, Pfrieger FW, Barres BA. Characterization of the signaling interactions that promote the survival and growth of developing retinal ganglion cells in culture. Neuron. 1995; 15:805–819. [PubMed: 7576630] Michaelis ML, Dobrowsky RT, Li G. Tau neurofibrillary pathology and microtubule stability. J Mol Neurosci. 2002; 19:289–293. [PubMed: 12540054] 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. 2000; 74:1008–1016. [PubMed: 10693931] Mizuno T, Nakata M, Naiki H, Michikawa M, Wang R, Haass C, Yanagisawa K. Cholesteroldependent generation of a seeding amyloid beta-protein in cell culture. J Biol Chem. 1999; 274:15110–15114. [PubMed: 10329717] Morishima-Kawashima M, Ihara Y. The presence of amyloid beta-protein in the detergent-insoluble membrane compartment of human neuroblastoma cells. Biochemistry. 1998; 37:15247–15253. [PubMed: 9799484] Nicholson AM, Ferreira A. Increased membrane cholesterol might render mature hippocampal neurons more susceptible to beta-amyloid-induced calpain activation and tau toxicity. J Neurosci. 2009; 29:4640–4651. [PubMed: 19357288] Oshima N, Morishima-Kawashima M, Yamaguchi H, Yoshimura M, Sugihara S, Khan K, Games D, Schenk D, Ihara Y. Accumulation of amyloid beta-protein in the low-density membrane domain accurately reflects the extent of beta-amyloid deposition in the brain. Am J Pathol. 2001; 158:2209–2218. [PubMed: 11395399] Osono Y, Woollett LA, Herz J, Dietschy JM. Role of the low density lipoprotein receptor in the flux of cholesterol through the plasma and across the tissues of the mouse. J Clin Invest. 1995; 95:1124– 1132. [PubMed: 7883961] Park SY, Ferreira A. The generation of a 17 kDa neurotoxic fragment: an alternative mechanism by which tau mediates beta-amyloid-induced neurodegeneration. J Neurosci. 2005; 25:5365–5375. [PubMed: 15930385] Park SY, Tournell C, Sinjoanu RC, Ferreira A. Caspase-3- and calpain-mediated tau cleavage are differentially prevented by estrogen and testosterone in beta-amyloid-treated hippocampal neurons. Neuroscience. 2007; 144:119–127. [PubMed: 17055174] Pericak-Vance MA, Bebout JL, Gaskell PC Jr, Yamaoka LH, Hung WY, Alberts MJ, Walker AP, Bartlett RJ, Haynes CA, Welsh KA, L EN, Heyman A, Clark CM, Roses AD. Linkage studies in familial Alzheimer disease: evidence for chromosome 19 linkage. Am J Hum Genet. 1991; 48:1034–1050. [PubMed: 2035524] Pfrieger FW. Cholesterol homeostasis and function in neurons of the central nervous system. Cell Mol Life Sci. 2003a; 60:1158–1171. [PubMed: 12861382] Pfrieger FW. Role of cholesterol in synapse formation and function. Biochim Biophys Acta. 2003b; 1610:271–280. [PubMed: 12648780] Poirier J. Apolipoprotein E in animal models of CNS injury and in Alzheimer's disease. Trends Neurosci. 1994; 17:525–530. [PubMed: 7532337] Poirier J, Baccichet A, Dea D, Gauthier S. Cholesterol synthesis and lipoprotein reuptake during synaptic remodelling in hippocampus in adult rats. Neuroscience. 1993; 55:81–90. [PubMed: 8350994] Raastad M, Shepherd GM. Single-axon action potentials in the rat hippocampal cortex. J Physiol. 2003; 548:745–752. [PubMed: 12640022] Racchi M, Baetta R, Salvietti N, Ianna P, Franceschini G, Paoletti R, Fumagalli R, Govoni S, Trabucchi M, Soma M. Secretory processing of amyloid precursor protein is inhibited by increase in cellular cholesterol content. Biochem J. 1997; 322(Pt 3):893–898. [PubMed: 9148766] Rebeck GW, Reiter JS, Strickland DK, Hyman BT. Apolipoprotein E in sporadic Alzheimer's disease: allelic variation and receptor interactions. Neuron. 1993; 11:575–580. [PubMed: 8398148]

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 13

NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Refolo LM, Malester B, LaFrancois J, Bryant-Thomas T, Wang R, Tint GS, Sambamurti K, Duff K, Pappolla MA. Hypercholesterolemia accelerates the Alzheimer's amyloid pathology in a transgenic mouse model. Neurobiol Dis. 2000; 7:321–331. [PubMed: 10964604] Rockwood K, Kirkland S, Hogan DB, MacKnight C, Merry H, Verreault R, Wolfson C, McDowell I. Use of lipid-lowering agents, indication bias, and the risk of dementia in community-dwelling elderly people. Arch Neurol. 2002; 59:223–227. [PubMed: 11843693] Roher AE, Lowenson JD, Clarke S, Wolkow C, Wang R, Cotter RJ, Reardon IM, Zurcher-Neely HA, Heinrikson RL, Ball MJ, et al. Structural alterations in the peptide backbone of beta-amyloid core protein may account for its deposition and stability in Alzheimer's disease. J Biol Chem. 1993; 268:3072–3083. [PubMed: 8428986] Saito M, Benson EP, Saito M, Rosenberg A. Metabolism of cholesterol and triacylglycerol in cultured chick neuronal cells, glial cells, and fibroblasts: accumulation of esterified cholesterol in serumfree culture. J Neurosci Res. 1987; 18:319–325. [PubMed: 3694714] Scott HD, Laake K. Statins for the prevention of Alzheimer's disease. Cochrane Database Syst Rev. 2001 CD003160. Shanmugaratnam J, Berg E, Kimerer L, Johnson RJ, Amaratunga A, Schreiber BM, Fine RE. Retinal Muller glia secrete apolipoproteins E and J which are efficiently assembled into lipoprotein particles. Brain Res Mol Brain Res. 1997; 50:113–120. [PubMed: 9406925] Shie FS, Jin LW, Cook DG, Leverenz JB, LeBoeuf RC. Diet-induced hypercholesterolemia enhances brain A beta accumulation in transgenic mice. Neuroreport. 2002; 13:455–459. [PubMed: 11930160] Simons K, Ikonen E. Functional rafts in cell membranes. Nature. 1997; 387:569–572. [PubMed: 9177342] Simons K, Toomre D. Lipid rafts and signal transduction. Nat Rev Mol Cell Biol. 2000; 1:31–39. [PubMed: 11413487] Simons K, Ikonen E. How cells handle cholesterol. Science. 2000; 290:1721–1726. [PubMed: 11099405] Simons M, Keller P, De Strooper B, Beyreuther K, Dotti CG, Simons K. Cholesterol depletion inhibits the generation of beta-amyloid in hippocampal neurons. Proc Natl Acad Sci U S A. 1998; 95:6460–6464. [PubMed: 9600988] Sisodia SS, Koo EH, Beyreuther K, Unterbeck A, Price DL. Evidence that beta-amyloid protein in Alzheimer's disease is not derived by normal processing. Science. 1990; 248:492–495. [PubMed: 1691865] Smith IF, Green KN, LaFerla FM. Calcium dysregulation in Alzheimer's disease: recent advances gained from genetically modified animals. Cell Calcium. 2005; 38:427–437. [PubMed: 16125228] Solomon A, Kivipelto M, Wolozin B, Zhou J, Whitmer RA. Midlife Serum Cholesterol and Increased Risk of Alzheimer's and Vascular Dementia Three Decades Later. Dement Geriatr Cogn Disord. 2009; 28:75–80. [PubMed: 19648749] Spady DK, Dietschy JM. Sterol synthesis in vivo in 18 tissues of the squirrel monkey, guinea pig, rabbit, hamster, and rat. J Lipid Res. 1983; 24:303–315. [PubMed: 6842086] Sparks DL, Liu H, Scheff SW, Coyne CM, Hunsaker JC 3rd. Temporal sequence of plaque formation in the cerebral cortex of non-demented individuals. J Neuropathol Exp Neurol. 1993; 52:135– 142. [PubMed: 8440995] Sparks DL, Hunsaker JC 3rd, Scheff SW, Kryscio RJ, Henson JL, Markesbery WR. Cortical senile plaques in coronary artery disease, aging and Alzheimer's disease. Neurobiol Aging. 1990; 11:601–607. [PubMed: 1704106] Sparks DL, Scheff SW, Hunsaker JC 3rd, Liu H, Landers T, Gross DR. Induction of Alzheimer-like beta-amyloid immunoreactivity in the brains of rabbits with dietary cholesterol. Exp Neurol. 1994; 126:88–94. [PubMed: 8157129] Sparks DL, Sabbagh MN, Connor DJ, Lopez J, Launer LJ, Browne P, Wasser D, Johnson-Traver S, Lochhead J, Ziolwolski C. Atorvastatin for the treatment of mild to moderate Alzheimer disease: preliminary results. Arch Neurol. 2005; 62:753–757. [PubMed: 15883262]

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

Nicholson and Ferreira

Page 14

NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Suzuki K, Parker CC, Pentchev PG, Katz D, Ghetti B, D'Agostino AN, Carstea ED. Neurofibrillary tangles in Niemann-Pick disease type C. Acta Neuropathol. 1995; 89:227–238. [PubMed: 7754743] Suzuki T. Lipid rafts at postsynaptic sites: distribution, function and linkage to postsynaptic density. Neurosci Res. 2002; 44:1–9. [PubMed: 12204288] Swanson LW, Simmons DM, Hofmann SL, Goldstein JL, Brown MS. Localization of mRNA for low density lipoprotein receptor and a cholesterol synthetic enzyme in rabbit nervous system by in situ hybridization. Proc Natl Acad Sci U S A. 1988; 85:9821–9825. [PubMed: 2462254] Tabernero A, Bolanos JP, Medina JM. Lipogenesis from lactate in rat neurons and astrocytes in primary culture. Biochem J. 1993; 294(Pt 3):635–638. [PubMed: 8379917] Takashima A. GSK-3 is essential in the pathogenesis of Alzheimer's disease. J Alzheimers Dis. 2006; 9:309–317. [PubMed: 16914869] Tokuda T, Tamaoka A, Matsuno S, Sakurai S, Shimada H, Morita H, Ikeda S. Plasma levels of amyloid beta proteins did not differ between subjects taking statins and those not taking statins. Ann Neurol. 2001; 49:546–547. [PubMed: 11310640] Torp R, Head E, Milgram NW, Hahn F, Ottersen OP, Cotman CW. Ultrastructural evidence of fibrillar beta-amyloid associated with neuronal membranes in behaviorally characterized aged dog brains. Neuroscience. 2000; 96:495–506. [PubMed: 10717430] Tsai LH, Lee MS, Cruz J. Cdk5, a therapeutic target for Alzheimer's disease? Biochim Biophys Acta. 2004; 1697:137–142. [PubMed: 15023356] Turley SD, Burns DK, Dietschy JM. Preferential utilization of newly synthesized cholesterol for brain growth in neonatal lambs. Am J Physiol. 1998; 274:E1099–E1105. [PubMed: 9611162] Urmoneit B, Reinsch C, Turner J, Czech C, Beyreuther K, Dyrks T. Inhibition of beta A4 production by specific modulation of beta-secretase activity. J Mol Neurosci. 1995; 6:23–32. [PubMed: 8562317] Vassar R, Bennett BD, Babu-Khan S, Kahn S, Mendiaz EA, Denis P, Teplow DB, Ross S, Amarante P, Loeloff R, Luo Y, Fisher S, Fuller J, Edenson S, Lile J, Jarosinski MA, Biere AL, Curran E, Burgess T, Louis JC, Collins F, Treanor J, Rogers G, Citron M. Beta-secretase cleavage of Alzheimer's amyloid precursor protein by the transmembrane aspartic protease BACE. Science. 1999; 286:735–741. [PubMed: 10531052] Volpe JJ, Hennessy SW. Cholesterol biosynthesis and 3-hydroxy-3-methyl-glutaryl coenzyme A reductase in cultured glial and neuronal cells. Regulation by lipoprotein and by certain free sterols. Biochim Biophys Acta. 1977; 486:408–420. [PubMed: 856284] Wolozin B, Kellman W, Ruosseau P, Celesia GG, Siegel G. Decreased prevalence of Alzheimer disease associated with 3-hydroxy-3-methyglutaryl coenzyme A reductase inhibitors. Arch Neurol. 2000; 57:1439–1443. [PubMed: 11030795] Wood JG, Mirra SS, Pollock NJ, Binder LI. Neurofibrillary tangles of Alzheimer disease share antigenic determinants with the axonal microtubule-associated protein tau (tau). Proc Natl Acad Sci U S A. 1986; 83:4040–4043. [PubMed: 2424015] Xia W, Zhang J, Perez R, Koo EH, Selkoe DJ. Interaction between amyloid precursor protein and presenilins in mammalian cells: implications for the pathogenesis of Alzheimer disease. Proc Natl Acad Sci U S A. 1997; 94:8208–8213. [PubMed: 9223340] Yamaguchi H, Maat-Schieman ML, van Duinen SG, Prins FA, Neeskens P, Natte R, Roos RA. Amyloid beta protein (Abeta) starts to deposit as plasma membrane-bound form in diffuse plaques of brains from hereditary cerebral hemorrhage with amyloidosis-Dutch type, Alzheimer disease and nondemented aged subjects. J Neuropathol Exp Neurol. 2000; 59:723–732. [PubMed: 10952062]

Cogn Sci (Hauppauge). Author manuscript; available in PMC 2014 October 20.

CHOLESTEROL AND NEURONAL SUSCEPTIBILITY TO BETA-AMYLOID TOXICITY.

Alzheimer's disease (AD) is a devastating neurocognitive disorder rapidly growing across the elderly population. Although few cases arise due to genet...
341KB Sizes 2 Downloads 6 Views