349
Views
0
CrossRef citations to date
0
Altmetric
Editorial

Do Prion-Induced Changes in Membrane Cholesterol Trigger Neurodegeneration?

&
Pages 367-370 | Published online: 12 Jun 2008

The transmissible spongiform encephalopathies (TSEs) are associated with the conversion of a host-encoded cellular protein (PrPC) to alternatively folded, disease-associated isoforms (PrPSc) Citation[1]. These diseases are unique in that the only identified component of the infectious agent is a self-replicating protein (PrPSc); which explains why such diseases are commonly called prion diseases (protein-only infectious agent). The accumulation of PrPSc is closely associated with the main pathological features of prion diseases; loss of synapses, activation of glial cells, neuronal loss and the spongiform degeneration of the brain Citation[2]. However, the cellular and molecular mechanisms leading to neurodegeneration in these diseases are imperfectly understood. In this review, we outline developments implicating prion-induced changes in the levels of membrane cholesterol as a trigger of abnormal cell signaling, which leads to neurodegeneration.

Recent studies have demonstrated that prion infection has a significant impact on the biochemistry of neuronal cell membranes; it significantly increased the amounts of free cholesterol, but reduced the amounts of esterified cholesterol Citation[3]. In addition, a highly significant correlation between the amounts of PrPSc and the concentration of free cholesterol in cell membranes was observed. Crucially, the effects of prion infection were not replicated by stimulating cholesterol biosynthesis. Stimulation of cholesterol production in uninfected neurons had the opposite effect to prion infection; it increased cholesterol ester formation without altering the amounts of free cholesterol. Such observations indicate that prion infection increased the capacity of cell membranes to solubilize free cholesterol. Cellular cholesterol is found as either free cholesterol in membranes, or as cholesterol esters that are stored in cytoplasmic droplets. A dynamic equilibrium between the pools of free cholesterol and cholesterol esters is tightly controlled by acyl-coenzyme A:cholesterol acyltransferase (ACAT), an endoplasmic reticulum (ER)-resident enzyme that esterifies free cholesterol with long-chain fatty acids Citation[4]. Therefore, the combination of increased free cholesterol and reduced cholesterol esters observed in prion-infected cells may result from inhibition of ACAT, or from the sequestration of free cholesterol in microenvironments that avoid ACAT.

The capacity of cell membranes to solubilize free cholesterol is partly determined by its fatty acid composition. The high incidence of saturated fatty acids attached to glycosylphosphatidylinositol (GPI)-anchored proteins allows tight molecular packing and increases the solubilization of free cholesterol Citation[5]. Since both PrPCand PrPScare linked to membranes via a GPI anchor Citation[6], the formation of PrPSc can have a direct effect on the composition of cell membranes. One property of PrPSc molecules that is not shared by PrPC is the propensity to self-aggregate. As a consequence, the formation of PrPScoligomers results in the clustering of GPI anchors in high densities. We propose that the increased density of saturated fatty acids within PrPSc-containing microdomains encourages the solubilization of free cholesterol. This hypothesis is consistent with observations that the addition of PrPSc with intact GPI anchors to uninfected neurons increased free cholesterol within cell membranes, while the addition of PrPSc with a GPI anchor lacking acyl chains had no effect. Similarly, the addition of synthetic GPI analogues containing two saturated acyl chains significantly increased the free cholesterol content of cell membranes, while GPI analogues containing either a single saturated acyl chain, or none at all, had no effect [Bate & Williams, Unpublished Data].

So what are the possible consequences of increased free cholesterol in cell membranes? Although free cholesterol is a component of normal cell membranes, the amounts of free cholesterol are increased up to fivefold in specialized microdomains within the plasma membrane that are commonly called lipid rafts Citation[7]. Moreover, the formation of some of these lipid rafts is sensitive to changes in membrane cholesterol concentrations. Since GPI-anchored proteins, including PrPC and PrPSc, are targeted to lipid rafts Citation[8], the formation of PrPSc may alter conventional raft structure and the PrPC–protein interactions that occur within lipid rafts. For example, PrPC has been reported to bind to caveolin-1 Citation[9] and neural cell adhesion molecule Citation[10], proteins that reside within lipid rafts. It is unclear if these protein–protein interactions are affected following the conversion of PrPC to PrPSc.

Currently, some lipid rafts are thought to act as membrane platforms in which signaling complexes assemble Citation[11]. Since the formation of some of these rafts is sensitive to fluctuations in membrane cholesterol concentrations, prion-induced changes in free cholesterol levels within cell membranes may have profound effects on cell signaling. The unregulated activation of phospholipase A2 (PLA2) is now recognized as a key event in neurodegenerative diseases Citation[12], including prion diseases Citation[13]. Not only did prion-infected cells contain four-times more activated cytoplasmic PLA2 (cPLA2) than uninfected cells, but the amounts of activated cPLA2 correlated with both the amounts of free cholesterol and with the amounts of PrPScCitation[3]. Immunoprecipitation studies demonstrated that cPLA2 colocalized with PrPSc-containing lipid rafts, suggesting that the capture of free cholesterol by PrPSc was required for cPLA2 activation. A causal relationship between cPLA2 and PrPSc is consistent with reports that PLA2 activation in prion-infected neuronal cells was cholesterol-sensitive Citation[14] and that inhibition of cholesterol esterification increased free cholesterol, increased activation of PLA2 and precipitated the death of prion-infected cells Citation[15]. The activation of PLA2 is the first step in the production of eicosanoids, docosanoids and platelet-activating factors, and although such molecules have a role in the maintenance of neuronal function, increased production of these factors causes synapse damage, glial cell activation and neuronal death. Although our studies focused on the PLA2 pathway, many other signaling molecules, including the Src family tyrosine kinases Citation[16], adenylyl cyclase Citation[17] and the trimeric G-proteins Citation[18], are also found in cholesterol-sensitive lipid rafts. The effect of prion-induced changes in free cholesterol levels on the activity of these signaling pathways is not currently known.

Abnormal cell signaling may not be the only pathological consequence of increased cholesterol. In neuronal cells, free cholesterol is readily oxidized into cytotoxic oxysterols Citation[19]. Thus, increased free cholesterol may enhance neuronal death under oxidative conditions. Changes in the cholesterol content of membranes may also affect the activity of the Hedgehog family of tissue-patterning factors that are covalently modified by cholesterol Citation[20]. Increasing the free cholesterol content of membranes may also affect membrane fluidity, endocytosis and intracellular trafficking of proteins Citation[21]. Another affect of the sequestration of free cholesterol into PrPSc-containing lipid rafts may be depletion of free cholesterol from other cellular pools where it helps to stabilize the packing of sphingolipids, gangliosides and raft-associated proteins in the membrane. Thus, the sequestration of free cholesterol by PrPSchas the potential to affect the function of many lipid raft-associated proteins. The receptors for acetylcholine, glutamate and γ-aminobutyric acid are found within lipid rafts and are affected by cellular cholesterol levels Citation[22]. Since free cholesterol is also required for the formation and function of synapses Citation[23], the sequestration of cholesterol by PrPSc may affect synaptic transmission. This hypothesis is supported by experimental data showing that prion-infected cells contain reduced amounts of synaptic proteins Citation[24] and that the loss of synapses is characteristic of prion diseases Citation[25].

Disturbances in membrane cholesterol are observed in other neurodegenerative conditions, including Alzheimer‘s and Parkinson‘s diseases Citation[21]. The accumulation of amyloid-β peptides in Alzheimer‘s disease, and α-synuclein in Parkinson‘s disease, within cholesterol-rich domains in the brain is thought to play a central role in neuropathogenesis Citation[26,27]. The similarities in the pathogenesis of such disorders and prion diseases raise the possibility that disturbed cholesterol metabolism may also contribute towards neurodegeneration in these diseases.

Much of what is known about the role of cholesterol in cell membranes is surmised from the changes in cells following cholesterol depletion. In particular, the statins, drugs that inhibit cholesterol biosynthesis, are increasingly being used to treat neurodegenerative conditions, including Alzheimer‘s disease Citation[28], Parkinson‘s disease Citation[29] and multiple sclerosis Citation[30]. However, considering the number of neuronal processes that require cholesterol, the use of statins can be viewed as a crude pharmacological tool. Moreover, it is now clear that lipid rafts exist as heterogeneous subsets, each with different compositions and different functions Citation[7]. These observations raise the intriguing possibility that drugs can be designed to alter the function of specific lipid raft subsets. Major future goals must surely include the development of novel therapeutics that are able to selectively target specific lipid rafts or raft functions.

We propose that the clustering of GPI anchors in prion diseases increases the solubilization of free cholesterol in specific membrane microdomains. This reduces the cycling of free cholesterol through the ER and exposure to ACAT, resulting in a reduction in cholesterol-ester formation. The increased amounts of free cholesterol in specific domains increase PLA2 activation, a key event in prion-induced neurotoxicity. We further propose that similar perturbations in membrane cholesterol levels and altered cell signaling, albeit caused by different agents, may be a common factor in neurodegenerative diseases. Novel compounds that selectively disrupt prion, amyloid-β or α-synuclein-induced raft formation and activity may constitute the next generation of therapeutics for neurodegenerative diseases.

Financial & competing interests disclosure

The authors have no relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript. This includes employment, consultancies, honoraria, stock ownership or options, expert testimony, grants or patents received or pending, or royalties.

No writing assistance was utilized in the production of this manuscript.

Bibliography

  • Prusiner SB : Prions.Proc. Natl Acad. Sci. USA95, 13363–13383 (1998).
  • Cunningham C , DeaconR, WellsHet al.: Synaptic changes characterize early behavioural signs in the ME7 model of murine prion disease. Eur. J. Neurosci.17, 2147–2155 (2003).
  • Bate C , TayebiM, WilliamsA: Sequestration of free cholesterol in cell membranes by prions correlates with cytoplasmic phospholipase A2 activation.BMC Biol.6, 8 (2008).
  • Chang T -Y, Chang CCY, Ohgami N, Yamauchi Y: Cholesterol sensing, trafficking, and esterification. Annu. Rev. Cell Dev. Biol.22, 129–157 (2006).
  • Schroeder F , GallegosAM, AtshavesMPet al.: Recent advances in membrane microdomains: rafts, caveolae, and intracellular cholesterol trafficking. Exp. Biol. Med226, 873–890 (2001).
  • Stahl N , BaldwinMA, HeckerR, PanKM, BurlingameAL, PrusinerSB: Glycosylinositol phospholipid anchors of the scrapie and cellular prion proteins contain sialic acid.Biochemistry31, 5043–5053 (1992).
  • Pike LJ : Lipid rafts: heterogeneity on the high seas.Biochem. J.378, 281–292 (2004).
  • Vey M , PilkuhnS, WilleHet al.: Subcellular colocalization of the cellular and scrapie prion proteins in caveolae-like membranous domains. Proc. Natl Acad. Sci. USA93, 14945–14949 (1996).
  • Massimino ML , GriffoniC, SpisniE, ToniM, TomasiV: Involvement of caveolae and caveolae-like domains in signalling, cell survival and angiogenesis.Cell. Signal.14, 93–98 (2002).
  • Schmitt-Ulms G , LegnameG, BaldwinMAet al.: Binding of neural cell adhesion molecules (N-CAMs) to the cellular prion protein. J. Mol. Biol.314, 1209–1225 (2001).
  • Tsui-Pierchala BA , EncinasM, MilbrandtJ, JohnsonEM: Lipid rafts in neuronal signaling and function.Trends Neurosci.25, 412–417 (2002).
  • Farooqui AA , Ong W-Y, Horrocks LA: Inhibitors of brain phospholipase A2 activity: their neuropharmacological effects and therapeutic importance for the treatment of neurologic disorders. Phamacol. Rev.58, 591–620 (2006).
  • Bate C , ReidS, WilliamsA: Phospholipase A2 inhibitors or platelet activating factor antagonists prevent prion replication.J. Biol. Chem.279, 36405–36411 (2004).
  • Bate C , SalmonaM, DiomedeL, WilliamsA: Squalestatin cures prion-infected neurones and protects against prion neurotoxicity.J. Biol. Chem.279, 14983–14990 (2004).
  • Bate C , TayebiM, WilliamsA: Cholesterol esterification reduces the neurotoxicity of prions.Neuropharmacology DOI: 10.1016/j.neuropharm.2008.04.002 (2008) (Epub ahead of print).
  • Li S , CouetJ, LisantiML: Src tyrosine kinases, Gα Subunits, and H-Ras share a common membrane-anchored scaffolding protein, caveolin. Caveolin binding negatively regulates the auto-activation of Src tyrosine kinases.J. Biol. Chem.271, 29182–29190 (1996).
  • Cooper DMF , CrossthwaiteAJ: Higher-order organization and regulation of adenylyl cyclases.Trends Pharmacol. Sci.27, 426–431 (2006).
  • Oh P , SchnitzerJE: Segregation of heterotrimeric G proteins in cell surface microdomains. Gq binds caveolin to concentrate in caveolae, whereas Gi and Gs target lipid rafts by default.Mol. Biol. Cell12, 685–698 (2001).
  • Sathishkumar K , MurthySN, UppuRM: Cytotoxic effects of oxysterols produced during ozonolysis of cholesterol in murine GT1–7 hypothalamic neurons.Free Radic. Res.41, 82–88 (2007).
  • Mann RK , BeachyPA: Cholesterol modification of proteins.Biochim. Biophys. Acta1529, 188–202 (2000).
  • Maxfield FR , TabasI: Role of cholesterol and lipid organization in disease.Nature438, 612–21 (2005).
  • Allen JA , Halverson-TamboliRA, Rasenick MM: Lipid raft microdomains and neurotransmitter signalling. Nat. Rev. Neurosci.8, 128–140 (2007).
  • Goritz C , MauchDH, PfriegerFW: Multiple mechanisms mediate cholesterol-induced synaptogenesis in a CNS neuron.Mol. Cell. Neurosci.29, 190–201 (2005).
  • Sandberg MK , LowP: Altered interaction and expression of proteins involved in neurosecretion in scrapie-infected GT1-1 cells.J. Biol. Chem.280, 1264–1271 (2005).
  • Ferrer I : Synaptic pathology and cell death in the cerebellum in Creutzfeldt–Jakob disease.Cerebellum1, 213–222 (2002).
  • Rottkamp CA , AtwoodCS, JosephJA, NunomuraA, PerryG, SmithMA: The state versus amyloid-β: the trial of the most wanted criminal in Alzheimer disease.Peptides23, 1333–1341 (2002).
  • Lee VMY , TrojanowskiJQ: Mechanisms of Parkinson‘s disease linked to pathological α-synuclein: new targets for drug discovery.Neuron52, 33–38 (2006).
  • Li G , LarsonEB, SonnenJAet al.: Statin therapy is associated with reduced neuropathologic changes of Alzheimer disease. Neurology69, 878–885 (2007).
  • Wolozin B , WangS, Li N-C, Lee A, Lee T, Kazis L: Simvastatin is associated with a reduced incidence of dementia and Parkinson‘s disease. BMC Med.5, 20 (2007).
  • Neuhaus O , StuveO, ZamvilSS, Hartung HP: Evaluation of HMG-CoA reductase inhibitors for multiple sclerosis: opportunities and obstacles. CNS Drugs19, 833–841 (2005).

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.