Skip to main content

Is cholesterol both the lock and key to abnormal transmembrane signals in Autism Spectrum Disorder?

Abstract

Disturbances in cholesterol homeostasis have been associated with ASD. Lipid rafts are central in many transmembrane signaling pathways (including mTOR) and changes in raft cholesterol content affect their order function. Cholesterol levels are controlled by several mechanisms, including endoplasmic reticulum associated degradation (ERAD) of the rate limiting HMGCoA reductase. A new approach to increase cholesterol via temporary ERAD blockade using a benign bacterial toxin-derived competitor for the ERAD translocon is suggested.

A new lock and key model for cholesterol/lipid raft dependent signaling is proposed in which the rafts provide both the afferent and efferent ‘tumblers’ across the membrane to allow ‘lock and key’ receptor transmembrane signals.

Introduction

Transmembrane signalling is key to cellular homeostasis but is, in particular, the central feature of neurotransmission. The concept of lipid rafts as more ordered, cholesterol enriched microdomains within plasma membranes, first proposed by Simons [1, 2], has put lipid organization at the forefront of membrane biology. While the heterogeneity and dynamics of cellular rafts remain controversial in part, it is clear that cholesterol, which distinguishes eukaryotic membranes, impinges all aspects of membrane signalling. The brain makes its own cholesterol and has the highest tissue content (-25% of the total), mostly in myelin. Cholesterol rafts play an important role in neurotransmission and brain diseases. This is most clear in autism where the disparate risks, including mTOR signaling, may residue under this umbrella.

Lipid raft composition

Membrane proteins are either associated with or dissociated from, lipid rafts [3]. Membrane proteins are less fluid in raft than non-raft membranes [4]. Lipid membrane anchors [5, 6] hydrophobic matching [7] and protein/protein interactions [8] effect protein partitioning into lipid raft. It is expected that signaling pathways directly involving membrane raft lipids [9, 10] would be more prone to a raft dependent mechanism of transduction (as proposed in Fig. 1). Cholesterol, responsible for their increased order (Lo phase [11], or rigidity) is preferentially found in the outer exoplasmic PM leaflet, but lipid rafts and (lower levels of) cholesterol are also found in the cytosolic membrane leaflet [12,13,14], although the cytoplasmic species are more unsaturated and less ordered [15]. These opposing leaflet domains may interact in a regulatory manner [16, 17].

The phospholipid species of the exoplasmic and cytoplasmic plasma membrane leaflets are distinct [15]. PC and sphingomyelin are the primary exoplasmic species, while PS and PE are restricted to the cytoplasmic leaflet. Sub-species of PE can interact with cholesterol to form ordered domains [18]. This phospholipid asymmetry, maintained by ABC transporters [19,20,21], which are in turn, regulated by lipid rafts [22], is one of the first casualties of apoptosis [23].

The planar hydrophobic surface of cholesterol promotes an association with long chain hydrophobic species, particularly GSLs with saturated acyl moieties [24]. The wide array of lipid moieties contained within GSLs regulates their membrane [25] and raft distribution [26, 27]. The interaction of cholesterol can affect the surface conformation of the carbohydrate of GSLs from a membrane perpendicular to membrane parallel format [28]. This can restrict the access to exogenous GSL binding ligands [29]. Depletion of cholesterol can greatly increase the membrane binding of such ligands [30]. In addition, the membrane parallel conformation of GSL carbohydrate can have an ‘umbrella’ effect and mask the hydrophilic(OH) surface of membrane cholesterol [28]. Ligand binding to membrane perpendicular GSL would alter the equilibrium between parallel and perpendicular GSL carbohydrate and could thus promote cholesterol access by reducing the umbrella effect. This in effect, represents a biological transistor. The regulation of membrane receptor clustering after ligand binding, via cholesterol enriched rafts is well studied [31, 32].

Membrane cholesterol enriched lipid rafts are required for neurological pathways with deficits associated with ASD

Liquid ordered lipid rafts are central to transmembrane signaling in cells [2, 33]. Cholesterol enrichment is the key membrane ordering component in lipid rafts [34, 35]—more rigid heterogeneous domains [36] in the plasma membrane [8] which can recruit, and are required by many transmembrane signaling molecules [37,38,39,40]. This includes the mTOR signaling pathway associated with ASD [41, 42], a pathway dependent on both cholesterol [43, 44] and lipid rafts [45], the primary focus of this compendium.

Many studies have implicated low cholesterol (especially ‘good’ cholesterol [46]) as an important factor in ASD [47,48,49,50,51,52]. Dietary cholesterol supplementation can provide an ASD therapeutic approach [53, 54]. Many diverse genetic factors have been described as risk factors for ASD, [55,56,57,58,59,60,61,62] which has made defining a common mechanism difficult. A disturbance in the cholesterol composition of plasma membrane lipid rafts however, could have a pleiotropic effect on neurological signaling pathways, linking otherwise unrelated signal transduction pathways. Recently it has been proposed that that Clostridial metabolites, which primarily inhibit cholesterol biosynthesis, are the primary cause of ASD [63]. Aberrant cholesterol metabolism may predict sensory ASD deficiencies [64]. Protein condensates/phase separation have been implicated as an additional basis for autism [65, 66] and cholesterol can modulate such condensates [67]. Coupling such condensates to lipid ordered membrane domains can define their function [68].

Cholesterol is important in nerve signal transduction [69] and neuronal survival [70]. Cholesterol dependent lipid rafts and receptor protein clustering (densification) [32] are central regulatory components of transmembrane signalling [4], transmembrane signalling is key to nerve synapse function [71] /neurotrophic receptor traffic [72] and synapse function can be defective in ASD [73]. This association has been the subject of excellent review [48, 51, 52].

The mechanism by which low cholesterol could impinge on ASD synaptic and neurotransmission deficiency could be pleiotropic, since many of the ASD associated genes involve cholesterol raft dependent signal transduction/trafficking pathways [49]. Recent studies have further emphasized this linkage. Lipid rafts play key roles in synapse plasticity [74, 75] and have been implicated with an increasing number of signal pathways genetically associated with ASD: mTOR [76, 77], dopamine transport [78, 79], contactin-associated protein-like 2 synapse protein [80], acetyl choline receptor [81] gamma amino butyric acid receptor [82] and glutamate synapses [75, 83,84,85] and downstream cytosolic reorganization [86]. Nerve membrane receptor clustering is required for synapse function [87]. PCSK9, a regulator of lipoprotein/cholesterol homeostasis and neuronal development/apoptosis [88] has been identified as an ASD risk [89]. Modulation of the lipid raft cholesterol composition is central in Fragile X patients [90] and the rat [91] and mouse model of this ASD [92] and correcting cholesterol levels in part, corrects these differences [92]. Defective neural cholesterol homeostasis is associated with ASD [93].

In addition, cholesterol binding proteins -containing the binding CARC sequence motif [94] found largely in proteins within the exoplasmic membrane leaflet, can mediate interaction with inner membrane proteins containing the mirror CRAC sequence [95], to further amplify the role of cholesterol in transbilayer interactions [96, 97]. Such interactions could similarly be modified by aberrant cholesterol levels.

Statins and ASD

High cholesterol is associated with coronary problems and oral administration of statins (inhibitors of HMGCoA reductase) is the standard clinical therapeutic stratagem. (These drugs also inhibit protein prenylation since the isoprenoid structures added post-translationally are derived from the same mevalonate pathway [98, 99]). Although the goal, is not to reduce cholesterol below normal and therapeutic efficacy may be limited [100], it is of relevance to consider whether statins have any cholesterol mediated effect on ASD (and other neurological disorders [101]). Moreover, aberrant cholesterol’s association with ASD could include higher levels (as in Rett syndrome [102]) which could also disturb raft dependent signaling. Statin inhibition of cholesterol synthesis can promote axon regeneration [103]. Significantly, statins affect mTOR signalling [43], strongly associated with ASD [42]. Lovastatin treatment of Fragile X rats [104] or Rett syndrome mice [105] prevented cognitive defects. In a double blind randomized, placebo controlled clinical trial in ASD children, simvastatin was found to have a significant beneficial effect monitored by physiological behavioural parameters [106]. In children with neurofibromatosis, a monogenic model for autism, simvastatin also effected brain areas associated with this pathology and showed improved behavioural response in 25% of patients [107]. Drug screening in a drosophila neurofibromatosis model, identified simvastatin as a potential treatment [108], but lovastatin, and not the more apolar simvastatin is effective in the mouse Fragile X model [109]. The mechanistic basis of these results is however, complicated by the dual action of statins on cholesterol and prenylation. Nevertheless prenylation can affect protein-lipid raft partitioning [5, 110] so the effects of reducing cholesterol and prenylation could be related.

Cholesterol homeostasis

The cellular control of cholesterol biosynthesis is complex, largely defined by a cholesterol sensing mechanism in the endoplasmic reticulum(ER). When cholesterol is low, the ER located transcription factor SREBP(sterol regulatory element binding protein) [111], regulating the transcription of genes required for cholesterol biosynthesis, is transported to the Golgi by SCAP(SREBP cleavage activating protein) [112] for activation by proteolytic cleavage and SREBP then transits to the nucleus to activate the cholesterol biosynthetic genes. Thus, low ER cholesterol stimulates cholesterol levels via gene transcription.

An additional ER cholesterol regulated pathway down regulates cholesterol levels by a post translational mechanism. In the cholesterol biosynthetic pathway, the enzyme, 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase is rate limiting, and one of the additional ways this enzyme levels are regulated is by an unusual process, that of endoplasmic reticulum associated degradation (ERAD). ERAD is the normal cellular quality control mechanism, eliminating nascent misfolded proteins during ER traffic via a chaperone mediated protein unfolding [113], and subsequent transit through the ER translocon (dislocon [114]) to the cytosol for proteosomal degradation, ensuring the dissemination of only correctly 3D folded protein. In a few cases however, the correctly folded protein is also subject to ERAD as a means of control, for example CFTR [115, 116] and HMG-CoA reductase [117]. In terms of HMG-CoA reductase, this pathway is activated via low ER sterol-induced binding to Insig, [118, 119], ubiquitinylation by associated ligases [118,119,120,121], to initiate HMGCoA reductase unfolding and ER translocon transit to the cytosolic proteosome for degradation, reducing its cellular expression and thereby, cholesterol biosynthesis [117]. Many ER stress protein mutations are related to ASD [122] which might also impinge such ER regulated cholesterol metabolism.

Novel ERAD-based means to address hypocholesterolemia

The ERAD pathway is hijacked by many microbial pathogens since it provides a means to access the cell cytosol from the lumen of the ER/Golgi endomembrane system [123]. These include several viruses [124, 125] and cholera and Shiga toxins [126,127,128]. These toxins enter the endomembrane system by means of their carbohydrate (glycolipid) pentameric B subunit cell surface receptor binding which initiates internalization and retrograde transport to the ER [129]. Here the B subunits separate from the catalytic A subunit. The A subunit contains an N-terminal peptide sequence that mimics an unfolded (misfolded) protein [130], which recruits the ERAD machinery to transmit the A subunit from the ER to the cytosol via the translocon [131], and by avoiding the proteasome, to refold and access its cytosolic target protein (adenylate cyclase for cholera toxin and ribosomal RNA for Shiga toxin). Because of this, cholera toxin has been often used as a tool to probe the mechanisms of ERAD [132,133,134].

Since the dimensions of the ER translocon accommodate only one protein at a time, we have used this toxin/ERAD hijack as a means to exogenously regulate ERAD [135]. Many genetic diseases are exacerbated by ERAD, in that gene mutations that do not completely inactivate protein function, nevertheless induce minor protein misfolding, and thence ERAD elimination to cause/exacerbate insufficiency disease symptoms [136]. Such genetic diseases include cystic fibrosis, Gauchers Disease, Tay Sachs Disease, Fabry Disease and many more [137]. By mutational inactivation of the toxin A subunit, we generated a benign tool (e.g. mutant cholera toxin- mCT in which the A subunit catalytic activity is removed and does not induce a stress response [138]). This can block (occupy/compete for) the ERAD translocon and thereby allow such partially misfolded but functional nascent mutant proteins (e.g. deltaF508CFTR chloride transporter in cystic fibrosis, N370S glucocerebrosidase in Gaucher) to escape degradation and function to ameliorate deficiency disease symptoms. This system works in cell disease models [135] and mCT is highly effective in a mouse model of CF(delta F508CFTR) to normalize chloride-dependent saliva production [up to > 2 × normal] ([139], a standard index of CFTR function. mCT (rather than other subunit toxins) is the preferred ERAD blockade since the cholera toxin receptor, GM1 ganglioside is expressed on virtually all mammalian cells.

Since normal cholesterol biosynthesis is regulated in part, by ERAD of HMGCoA-reductase, our mCT ERAD blockade approach [135] also offers a potential benign, titratable means to temporarily reduce HMG-CoA reductase degradation to increase cholesterol biosynthesis during hypocholesterolemia. Indeed, blockade of HMGCoA-reductase ERAD has already been shown to increase cholesterol [121]. Furthermore, cholera toxin is able to transit the blood brain barrier [140, 141] and could therefore also modulate neural cholesterol metabolism.

Lock and Key receptor binding- only half the mechanism

The concept that protein ligands bind to their membrane receptors by a lock and key molecular complementary mechanism, is well entrenched [142] and validated in molecular biochemistry [143] particularly enzyme mechanisms [144]. However, this is only (less than?) half the living picture. Lock and key essentially only provides insight into the control of ligand/ receptor binding. Such membrane receptors are often recruited to lipid rafts which is essential to their subsequent signal function [145, 146]. The question of how a signal is transmitted is not addressed. The concept of a conformational change is handwaving. Why do downstream transmembrane enzymes etc. become activated, cluster, change? The rest of the lock needs to be considered: the tumblers and escape mechanism.

The eukaryotic membrane is amazingly complex, particularly in its lipid content [15, 147]. Why are so many long chain isoforms made? Cholesterol distinguishes eukaryotes and lipid rafts have revolutionized the way we consider transmembrane signaling. Rafts are more rigid domains in the outer leaflet of the plasma membrane, primarily as a result of their increased cholesterol content [148]. Glycosphingolipids (GSLs) are also key components and the binding of cholera toxin to its glycosphingolipid receptor, GM1 ganglioside, has long been used as a cytochemical marker of lipid rafts [129, 149].

Lipid rafts are heterogeneous [36] but generate a platform which could provide the ‘tumblers’ which determine on or off, cluster or separate, associate or disassociate. What if any, is the rotatory component? Although it took man to invent the wheel, molecular rotation is well described in the mitochondrial and other ATPases as the proton pump mechanism to generate ATP [150] and hence life. However, molecular rotation as a control mechanism in transmembrane signalling has not been considered. When the key opens the lock, are there tumblers? do the tumblers turn? Which way? How far? With whom? In what plane? Is it energy dependent? What are they? A plausible scenario is shown in Fig. 1. Here we propose lipid rafts are the tumblers. The lipid raft ordered domains are considered cogs (delimited by line tension [151]) on either side of the plasma membrane [11] which can rotate around the ligand/receptor complex. In this model, ligand membrane receptor binding engages the aforementioned tumblers for transmembrane coordination of these cogs. In the lock and key schematic (right), the proximal tumbler (red) is the exoplasmic PM raft in which the receptor is embedded, while the distal is a cytosolic leaflet lipid raft(blue). The ligand binding mediated coordination of these interplanar rafts induces/amplifies signaling. The species within lipid rafts have restricted translational(lateral) freedom [152, 153] which would aid cohesive rotary lipid raft tumbler signal transmission. The interlocking of exoplasmic and cytoplasmic lipid rafts could be further regulated via additional lateral raft association in the upper or lower bilayer leaflet (Fig. 1). Such lateral raft interactions can serve to recruit the additional downstream components of the signaling pathway. Transmembrane protein receptors could function as the lock to mediate these interplanar raft interactions, while aligning lipid interdigitation [154] could prove a mechanism for peripheral membrane protein raft receptors, GPI anchored proteins [155], or indeed, the interdigitation of raft lipids themselves [10, 154, 156]. Very long chain fatty acid synthesis is considered to facilitate interleaflet lipid interdigitation [157, 158]. The synthesis of such fatty acids has recently been found to be essential in neuronal growth cone lipid rafts and cell polarity [159]. The raft cholesterol content is different between exoplasmic and cytoplasmic membrane lipid rafts [160] and will play a marked role in their “tumbler” function, affected by sterol deficiency (or excess).

Fig. 1
figure 1

Receptor-ligand: lock-key. Membrane receptor species within lipid rafts may effect transmembrane signaling following ligand binding, via lateral raft-mediated modulation of their membrane environment. Rotation for example, may impact adjacent exoplasmic domains to provide a code for transmembrane domain interactions and subsequent cytosolic signal propagation

To extend the lock and key simile, the cell could be considered as a “safe” with an extremely complex but interrelated membrane “combination”. Cholesterol provides the blueprint for the lipid raft tumblers to enable the various “keys” to access this safe.

Conclusions

Membrane order is an important player in transmembrane signalling and cholesterol plays a large part in determining membrane order via the dynamic formation of lipid rafts. These rafts can communicate from one membrane side to the other, and this linkage can mediate ligand-membrane receptor binding dependent signal transduction. The defects in cholesterol homeostasis in ASD (and other neurological diseases) suggests this role is particularly important in neural physiology and networks, and increasing cholesterol provides a target for intervention. We suggest a novel mechanism to achieve this increase and propose an addition to the ‘lock and key ‘concept for membrane receptor binding in which cholesterol lipid rafts provide the tumblers to allow and discriminate signals across the membrane.

Availability of data and materials

No datasets were generated or analysed during the current study.

References

  1. van Meer G, Simons K. Lipid polarity and sorting in epithelial cells. J Cell Biochem. 1988;36(1):51–8.

    Article  PubMed  Google Scholar 

  2. Harder T, Simons K. Caveolae, DIGs, and the dynamics of sphingolipid-cholesterol microdomains. Curr Opin Cell Biol. 1997;9(4):534–42.

    Article  CAS  PubMed  Google Scholar 

  3. Lebreton S, Paladino S, Zurzolo C. Selective roles for cholesterol and actin in compartmentalization of different proteins in the Golgi and plasma membrane of polarized cells. J Biol Chem. 2008;283(43):29545–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Lingwood D, Simons K. Lipid rafts as a membrane-organizing principle. Science. 2010;327(5961):46–50.

    Article  CAS  PubMed  Google Scholar 

  5. Melkonian KA, Ostermeyer AG, Chen JZ, Roth MG, Brown DA. Role of lipid modifications in targeting proteins to detergent-resistant membrane rafts. Many raft proteins are acylated, while few are prenylated. J Biol Chem. 1999;274(6):3910–7.

    Article  CAS  PubMed  Google Scholar 

  6. Pinaud F, Michalet X, Iyer G, Margeat E, Moore HP, Weiss S. Dynamic partitioning of a glycosyl-phosphatidylinositol-anchored protein in glycosphingolipid-rich microdomains imaged by single-quantum dot tracking. Traffic. 2009;10(6):691–712.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. McIntosh TJ, Vidal A, Simon SA. Sorting of lipids and transmembrane peptides between detergent-soluble bilayers and detergent-resistant rafts. Biophys J. 2003;85(3):1656–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Nicolau DV Jr, Burrage K, Parton RG, Hancock JF. Identifying optimal lipid raft characteristics required to promote nanoscale protein-protein interactions on the plasma membrane. Mol Cell Biol. 2006;26(1):313–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Myeong J, Park CG, Suh BC, Hille B. Compartmentalization of phosphatidylinositol 4,5-bisphosphate metabolism into plasma membrane liquid-ordered/raft domains. Proc Natl Acad Sci U S A. 2021;118(9).

  10. Li S, Huang F, Xia T, Shi Y, Yue T. Phosphatidylinositol 4,5-Bisphosphate Sensing Lipid Raft via Inter-Leaflet Coupling Regulated by Acyl Chain Length of Sphingomyelin. Langmuir. 2023;39(17):5995–6005.

    Article  CAS  PubMed  Google Scholar 

  11. Allender DW, Schick M. A theoretical basis for nanodomains. J Membr Biol. 2022;255(4–5):451–60.

    Article  CAS  PubMed  Google Scholar 

  12. Prior IA, Muncke C, Parton RG, Hancock JF. Direct visualization of Ras proteins in spatially distinct cell surface microdomains. J Cell Biol. 2003;160(2):165–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Hayashi M, Shimada Y, Inomata M, Ohno-Iwashita Y. Detection of cholesterol-rich microdomains in the inner leaflet of the plasma membrane. Biochem Biophys Res Commun. 2006;351(3):713–8.

    Article  CAS  PubMed  Google Scholar 

  14. Buwaneka P, Ralko A, Liu SL, Cho W. Evaluation of the available cholesterol concentration in the inner leaflet of the plasma membrane of mammalian cells. J Lipid Res. 2021;62:100084.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Lorent JH, Levental KR, Ganesan L, Rivera-Longsworth G, Sezgin E, Doktorova M, et al. Plasma membranes are asymmetric in lipid unsaturation, packing and protein shape. Nat Chem Biol. 2020;16(6):644–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Lin X, Zhang S, Ding H, Levental I, Gorfe AA. The aliphatic chain of cholesterol modulates bilayer interleaflet coupling and domain registration. FEBS Lett. 2016;590(19):3368–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. St Clair JW, Kakuda S, London E. Induction of ordered lipid raft domain formation by loss of lipid asymmetry. Biophys J. 2020;119(3):483–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Grzybek M, Kubiak J, Lach A, Przybylo M, Sikorski AF. A raft-associated species of phosphatidylethanolamine interacts with cholesterol comparably to sphingomyelin. A Langmuir-Blodgett monolayer study. PLoS ONE. 2009;4(3):e5053.

    Article  PubMed  PubMed Central  Google Scholar 

  19. Plosch T, Kosters A, Groen AK, Kuipers F. The ABC of hepatic and intestinal cholesterol transport. Handb Exp Pharmacol. 2005;170:465–82.

    Article  Google Scholar 

  20. Quazi F, Molday RS. Lipid transport by mammalian ABC proteins. Essays Biochem. 2011;50(1):265–90.

    CAS  PubMed  Google Scholar 

  21. Coleman JA, Quazi F, Molday RS. Mammalian P4-ATPases and ABC transporters and their role in phospholipid transport. Biochim Biophys Acta. 2013;1831(3):555–74.

    Article  CAS  PubMed  Google Scholar 

  22. Aye IL, Singh AT, Keelan JA. Transport of lipids by ABC proteins: interactions and implications for cellular toxicity, viability and function. Chem Biol Interact. 2009;180(3):327–39.

    Article  CAS  PubMed  Google Scholar 

  23. Mapes J, Chen YZ, Kim A, Mitani S, Kang BH, Xue D. CED-1, CED-7, and TTR-52 regulate surface phosphatidylserine expression on apoptotic and phagocytic cells. Curr Biol. 2012;22(14):1267–75.

    Article  CAS  PubMed  Google Scholar 

  24. Kinnun JJ, Bolmatov D, Lavrentovich MO, Katsaras J. Lateral heterogeneity and domain formation in cellular membranes. Chem Phys Lipids. 2020;232:104976.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Schmieder SS, Tatituri R, Anderson M, Kelly K, Lencer WI. Structural basis for acyl chain control over glycosphingolipid sorting and vesicular trafficking. Cell Rep. 2022;40(2):111063.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Wang TY, Silvius JR. Different sphingolipids show differential partitioning into sphingolipid/cholesterol-rich domains in lipid bilayers. Biophys J. 2000;79(3):1478–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Xu X, Bittman R, Duportail G, Heissler D, Vilcheze C, London E. Effect of the structure of natural sterols and sphingolipids on the formation of ordered sphingolipid/sterol domains (rafts). Comparison of cholesterol to plant, fungal, and disease-associated sterols and comparison of sphingomyelin, cerebrosides, and ceramide. J Biol Chem. 2001;276(36):33540–6.

    Article  CAS  PubMed  Google Scholar 

  28. Lingwood D, Binnington B, Róg T, Vattulainen I, Grzybek M, Coskun U, et al. Cholesterol modulates glycolipid conformation and receptor activity. Nature Chem Biol. 2011;7:260–2.

    Article  CAS  Google Scholar 

  29. Mahfoud R, Manis A, Binnington B, Ackerley C, Lingwood CA. A major fraction of glycosphingolipids in model and cellular cholesterol containing membranes are undetectable by their binding proteins. J Biol Chem. 2010;285(46):36049–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Novak A, Binnington B, Ngan B, Chadwick K, Fleshner N, Lingwood CA. Cholesterol masking membrane glycosphingolipid tumor-associated antigens reduces their immunodetection in human cancer biopsies. Glycobiology. 2013;23(11):1230–9.

    Article  CAS  PubMed  Google Scholar 

  31. Viola A, Schroeder S, Sakakibara Y, Lanzavecchia A. T lymphocyte costimulation mediated by reorganization of membrane microdomains. Science. 1999;283(5402):680–2.

    Article  CAS  PubMed  Google Scholar 

  32. Bernard C, Carotenuto AR, Pugno NM, Fraldi M, Deseri L. Modelling lipid rafts formation through chemo-mechanical interplay triggered by receptor-ligand binding. Biomech Model Mechanobiol. 2023.

  33. Isik OA, Cizmecioglu O. Rafting on the plasma membrane: lipid rafts in signaling and disease. Adv Exp Med Biol. 2023;1436:87–108.

    Article  CAS  PubMed  Google Scholar 

  34. Ouweneel AB, Thomas MJ, Sorci-Thomas MG. The ins and outs of lipid rafts: functions in intracellular cholesterol homeostasis, microparticles, and cell membranes: Thematic Review Series: Biology of Lipid Rafts. J Lipid Res. 2020;61(5):676–86.

    Article  CAS  PubMed  Google Scholar 

  35. Shi Y, Ruan H, Xu Y, Zou C. Cholesterol, eukaryotic lipid domains, and an evolutionary perspective of transmembrane signaling. Cold Spring Harb Perspect Biol. 2023;15(11).

  36. de Almeida RF, Loura LM, Fedorov A, Prieto M. Lipid rafts have different sizes depending on membrane composition: a time-resolved fluorescence resonance energy transfer study. J Mol Biol. 2005;346(4):1109–20.

    Article  PubMed  Google Scholar 

  37. Freeman MR, Cinar B, Kim J, Mukhopadhyay NK, Di Vizio D, Adam RM, et al. Transit of hormonal and EGF receptor-dependent signals through cholesterol-rich membranes. Steroids. 2007;72(2):210–7.

    Article  CAS  PubMed  Google Scholar 

  38. Bieberich E. It’s a lipid’s world: bioactive lipid metabolism and signaling in neural stem cell differentiation. Neurochem Res. 2012;37(6):1208–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Li S, Leshchyns’ka I, Chernyshova Y, Schachner M, Sytnyk V. The neural cell adhesion molecule (NCAM) associates with and signals through p21-activated kinase 1 (Pak1). J Neurosci. 2013;33(2):790–803.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Li B, Qin Y, Yu X, Xu X, Yu W. Lipid raft involvement in signal transduction in cancer cell survival, cell death and metastasis. Cell Prolif. 2022;55(1):e13167.

    Article  CAS  PubMed  Google Scholar 

  41. Sato A. mTOR, a potential target to treat autism spectrum disorder. CNS Neurol Disord Drug Targets. 2016;15(5):533–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Thomas SD, Jha NK, Ojha S, Sadek B. mTOR signaling disruption and its association with the development of autism spectrum disorder. Molecules. 2023;28(4).

  43. Wang Y, You S, Su S, Yeon A, Lo EM, Kim S, et al. Cholesterol-lowering intervention decreases mTOR complex 2 signaling and enhances antitumor immunity. Clin Cancer Res. 2022;28(2):414–24.

    Article  CAS  PubMed  Google Scholar 

  44. Xu J, Dang Y, Ren YR, Liu JO. Cholesterol trafficking is required for mTOR activation in endothelial cells. Proc Natl Acad Sci U S A. 2010;107(10):4764–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Hsu JL, Leu WJ, Hsu LC, Liu SP, Zhong NS, Guh JH. Para-Toluenesulfonamide Induces Anti-tumor Activity Through Akt-Dependent and -Independent mTOR/p70S6K Pathway: Roles of Lipid Raft and Cholesterol Contents. Front Pharmacol. 2018;9:1223.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Tierney E, Remaley AT, Thurm A, Jager LR, Wassif CA, Kratz LE, et al. Sterol and lipid analyses identifies hypolipidemia and apolipoprotein disorders in autism associated with adaptive functioning deficits. Transl Psychiatry. 2021;11(1):471.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Tierney E, Bukelis I, Thompson RE, Ahmed K, Aneja A, Kratz L, et al. Abnormalities of cholesterol metabolism in autism spectrum disorders. Am J Med Genet B Neuropsychiatr Genet. 2006;141B(6):666–8.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Wang H. Lipid rafts: a signaling platform linking cholesterol metabolism to synaptic deficits in autism spectrum disorders. Front Behav Neurosci. 2014;8:104.

    Article  PubMed  PubMed Central  Google Scholar 

  49. Lingwood C. Is cholesterol the key factor for autism? Am J Biomed Sci Res. 2020;7(6):483–6.

    Article  Google Scholar 

  50. Esposito CM, Buoli M, Ciappolino V, Agostoni C, Brambilla P. The role of cholesterol and fatty acids in the etiology and diagnosis of autism spectrum disorders. Int J Mol Sci. 2021;22(7).

  51. Lin J, de Rezende VL, de Aguiar da Costa M, de Oliveira J, Goncalves CL. Cholesterol metabolism pathway in autism spectrum disorder: From animal models to clinical observations. Pharmacol Biochem Behav. 2023;223:173522.

  52. Benachenhou S, Laroui A, Dionne O, Rojas D, Toupin A, Caku A. Cholesterol alterations in fragile X syndrome, autism spectrum disorders and other neurodevelopmental disorders. Int Rev Neurobiol. 2023;173:115–39.

    Article  PubMed  Google Scholar 

  53. Aneja A, Tierney E. Autism: the role of cholesterol in treatment. Int Rev Psychiatry. 2008;20(2):165–70.

    Article  PubMed  Google Scholar 

  54. Thurm A, Tierney E, Farmer C, Albert P, Joseph L, Swedo S, et al. Development, behavior, and biomarker characterization of Smith-Lemli-Opitz syndrome: an update. J Neurodev Disord. 2016;8:12.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Woodbury-Smith M, Scherer SW. Progress in the genetics of autism spectrum disorder. Dev Med Child Neurol. 2018;60(5):445–51.

    Article  PubMed  Google Scholar 

  56. Bay H, Haghighatfard A, Karimipour M, Seyedena SY, Hashemi M. Expression alteration of Neuroligin family gene in attention deficit and hyperactivity disorder and autism spectrum disorder. Res Dev Disabil. 2023;139:104558.

    Article  PubMed  Google Scholar 

  57. Dingemans AJM, Truijen KMG, van de Ven S, Bernier R, Bongers E, Bouman A, et al. The phenotypic spectrum and genotype-phenotype correlations in 106 patients with variants in major autism gene CHD8. Transl Psychiatry. 2022;12(1):421.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Kwon SJ, Hong KW, Choi S, Hong JS, Kim JW, Kim JW, et al. Association of 3-hydroxy-3-methylglutaryl-coenzyme A reductase gene polymorphism with obesity and lipid metabolism in children and adolescents with autism spectrum disorder. Metab Brain Dis. 2022;37(2):319–28.

    Article  CAS  PubMed  Google Scholar 

  59. Ma L, Yuan T, Li W, Guo L, Zhu D, Wang Z, et al. Dynamic functional connectivity alterations and their associated gene expression pattern in autism spectrum disorders. Front Neurosci. 2021;15:794151.

    Article  PubMed  Google Scholar 

  60. Nicotera AG, Amore G, Saia MC, Vinci M, Musumeci A, Chiavetta V, et al. Fibroblast Growth Factor Receptor 2 (FGFR2), a New Gene Involved in the Genesis of Autism Spectrum Disorder. Neuromolecular Med. 2023.

  61. Shi X, Lu C, Corman A, Nikish A, Zhou Y, Platt RJ, et al. Heterozygous deletion of the autism-associated gene CHD8 impairs synaptic function through widespread changes in gene expression and chromatin compaction. Am J Hum Genet. 2023;110(10):1750–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Wang N, Lv L, Huang X, Shi M, Dai Y, Wei Y, et al. Gene editing in monogenic autism spectrum disorder: animal models and gene therapies. Front Mol Neurosci. 2022;15:1043018.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Shaw W. Inhibition of the Beta-oxidation Pathway of Fatty Acids and Dopamine- Beta-hydroxylase by Phenyl Derivatives of Short- Chain Fatty Acids from Gastrointestinal Clostridia Bacteria is a (the) Major Cause of Autism. Integr Med (Encinitas). 2023;22(2):18–25.

    PubMed  Google Scholar 

  64. Ma Z, Xu L, Li Q, Li X, Shi Y, Zhang X, et al. Prediction model for sensory perception abnormality in autism spectrum disorder. Int J Mol Sci. 2023;24(3):2367.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Tsang B, Pritisanac I, Scherer SW, Moses AM, Forman-Kay JD. Phase separation as a missing mechanism for interpretation of disease mutations. Cell. 2020;183(7):1742–56.

    Article  CAS  PubMed  Google Scholar 

  66. Shih PY, Fang YL, Shankar S, Lee SP, Hu HT, Chen H, et al. Phase separation and zinc-induced transition modulate synaptic distribution and association of autism-linked CTTNBP2 and SHANK3. Nat Commun. 2022;13(1):2664.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Videv P, Mladenova K, Andreeva TD, Park JH, Moskova-Doumanova V, Petrova SD, et al. Cholesterol alters the phase separation in model membranes containing hBest1. Molecules. 2022;27(13).

  68. Wang HY, Chan SH, Dey S, Castello-Serrano I, Rosen MK, Ditlev JA, et al. Coupling of protein condensates to ordered lipid domains determines functional membrane organization. Sci Adv. 2023;9(17):eadf6205.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Korinek M, Gonzalez-Gonzalez IM, Smejkalova T, Hajdukovic D, Skrenkova K, Krusek J, et al. Cholesterol modulates presynaptic and postsynaptic properties of excitatory synaptic transmission. Sci Rep. 2020;10(1):12651.

    Article  PubMed  PubMed Central  Google Scholar 

  70. Krivoi, II, Petrov AM. Cholesterol and the safety factor for neuromuscular transmission. Int J Mol Sci. 2019;20(5).

  71. Boda B, Dubos A, Muller D. Signaling mechanisms regulating synapse formation and function in mental retardation. Curr Opin Neurobiol. 2010;20(4):519–27.

    Article  CAS  PubMed  Google Scholar 

  72. Kwan HR, Chan ZC, Bi X, Kutkowska J, Proszynski TJ, Chan CB, et al. Nerve-independent formation of membrane infoldings at topologically complex postsynaptic apparatus by caveolin-3. Sci Adv. 2023;9(24):eadg0183.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Ebrahimi-Fakhari D, Sahin M. Autism and the synapse: emerging mechanisms and mechanism-based therapies. Curr Opin Neurol. 2015;28(2):91–102.

    Article  CAS  PubMed  Google Scholar 

  74. Yang XL, Xiong WC, Mei L. Lipid rafts in neuregulin signaling at synapses. Life Sci. 2004;75(21):2495–504.

    Article  CAS  PubMed  Google Scholar 

  75. Guirland C, Suzuki S, Kojima M, Lu B, Zheng JQ. Lipid rafts mediate chemotropic guidance of nerve growth cones. Neuron. 2004;42(1):51–62.

    Article  CAS  PubMed  Google Scholar 

  76. Sural-Fehr T, Singh H, Cantuti-Catelvetri L, Zhu H, Marshall MS, Rebiai R, et al. Inhibition of the IGF-1-PI3K-Akt-mTORC2 pathway in lipid rafts increases neuronal vulnerability in a genetic lysosomal glycosphingolipidosis. Dis Model Mech. 2019;12(5).

  77. Zhao Q, Su H, Jiang W, Luo H, Pan L, Liu Y, et al. IGF-1 combined with OPN promotes neuronal axon growth in vitro through the IGF-1R/Akt/mTOR signaling pathway in lipid rafts. Neurochem Res. 2023;48(10):3190–201.

    Article  CAS  PubMed  Google Scholar 

  78. Nguyen M, Roth A, Kyzar EJ, Poudel MK, Wong K, Stewart AM, et al. Decoding the contribution of dopaminergic genes and pathways to autism spectrum disorder (ASD). Neurochem Int. 2014;66:15–26.

    Article  CAS  PubMed  Google Scholar 

  79. Kovtun O, Sakrikar D, Tomlinson ID, Chang JC, Arzeta-Ferrer X, Blakely RD, et al. Single-quantum-dot tracking reveals altered membrane dynamics of an attention-deficit/hyperactivity-disorder-derived dopamine transporter coding variant. ACS Chem Neurosci. 2015;6(4):526–34.

    Article  CAS  PubMed  Google Scholar 

  80. Chen N, Koopmans F, Gordon A, Paliukhovich I, Klaassen RV, van der Schors RC, et al. Interaction proteomics of canonical Caspr2 (CNTNAP2) reveals the presence of two Caspr2 isoforms with overlapping interactomes. Biochim Biophys Acta. 2015;1854(7):827–33.

    Article  CAS  PubMed  Google Scholar 

  81. Valles AS, Barrantes FJ. Dysregulation of neuronal nicotinic acetylcholine receptor-cholesterol crosstalk in autism spectrum disorder. Front Mol Neurosci. 2021;14:744597.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Wang X, Zhao Z, Guo J, Mei D, Duan Y, Zhang Y, et al. GABA(B1) receptor knockdown in prefrontal cortex induces behavioral aberrations associated with autism spectrum disorder in mice. Brain Res Bull. 2023;202:110755.

    Article  CAS  PubMed  Google Scholar 

  83. Mlinac-Jerkovic K, Ilic K, Zjalic M, Mandic D, Debeljak Z, Balog M, et al. Who’s in, who’s out? Re-evaluation of lipid raft residents. J Neurochem. 2021;158(3):657–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Swanwick CC, Shapiro ME, Vicini S, Wenthold RJ. Flotillin-1 promotes formation of glutamatergic synapses in hippocampal neurons. Dev Neurobiol. 2010;70(13):875–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Yao M, Meng M, Yang X, Wang S, Zhang H, Zhang F, et al. POSH regulates assembly of the NMDAR/PSD-95/Shank complex and synaptic function. Cell Rep. 2022;39(1):110642.

    Article  CAS  PubMed  Google Scholar 

  86. Sadybekov A, Tian C, Arnesano C, Katritch V, Herring BE. An autism spectrum disorder-related de novo mutation hotspot discovered in the GEF1 domain of Trio. Nat Commun. 2017;8(1):601.

    Article  PubMed  PubMed Central  Google Scholar 

  87. Miyazaki Y, Otsuka T, Yamagata Y, Endo T, Sanbo M, Sano H, et al. Oligodendrocyte-derived LGI3 and its receptor ADAM23 organize juxtaparanodal Kv1 channel clustering for short-term synaptic plasticity. Cell Rep. 2024;43(1):113634.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Wei H, Alberts I, Li X. The apoptotic perspective of autism. Int J Dev Neurosci. 2014;36:13–8.

    Article  PubMed  Google Scholar 

  89. Salem S, Mosaad R, Lotfy R, Ashaat E, Ismail S. PCSK9 involvement in autism etiology: sequence variations, protein concentration, and promoter methylation. Arch Med Res. 2023;54(6):102860.

    Article  CAS  PubMed  Google Scholar 

  90. Toupin A, Benachenhou S, Abolghasemi A, Laroui A, Galarneau L, Fulop T, et al. Association of lipid rafts cholesterol with clinical profile in fragile X syndrome. Sci Rep. 2022;12(1):2936.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Cartocci V, Catallo M, Tempestilli M, Segatto M, Pfrieger FW, Bronzuoli MR, et al. Altered brain cholesterol/isoprenoid metabolism in a rat model of autism spectrum disorders. Neuroscience. 2018;372:27–37.

    Article  CAS  PubMed  Google Scholar 

  92. Kalinowska M, Castillo C, Francesconi A. Quantitative profiling of brain lipid raft proteome in a mouse model of fragile X syndrome. PLoS ONE. 2015;10(4):e0121464.

    Article  PubMed  PubMed Central  Google Scholar 

  93. Segatto M, Tonini C, Pfrieger FW, Trezza V, Pallottini V. Loss of mevalonate/cholesterol homeostasis in the brain: a focus on autism spectrum disorder and Rett syndrome. Int J Mol Sci. 2019;20(13).

  94. Fantini J, Barrantes FJ. How cholesterol interacts with membrane proteins: an exploration of cholesterol-binding sites including CRAC, CARC, and tilted domains. Front Physiol. 2013;4:31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Fantini J, Di Scala C, Evans LS, Williamson PT, Barrantes FJ. A mirror code for protein-cholesterol interactions in the two leaflets of biological membranes. Sci Rep. 2016;6:21907.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Fantini J, Epand RM, Barrantes FJ. Cholesterol-recognition motifs in membrane proteins. Adv Exp Med Biol. 2019;1135:3–25.

    Article  CAS  PubMed  Google Scholar 

  97. Sarkar P, Chattopadhyay A. Cholesterol interaction motifs in G protein-coupled receptors: slippery hot spots? Wiley Interdiscip Rev Syst Biol Med. 2020;12(4):e1481.

    Article  CAS  PubMed  Google Scholar 

  98. Fenton JW 2nd, Jeske WP, Catalfamo JL, Brezniak DV, Moon DG, Shen GX. Statin drugs and dietary isoprenoids downregulate protein prenylation in signal transduction and are antithrombotic and prothrombolytic agents. Biochemistry (Mosc). 2002;67(1):85–91.

    Article  CAS  PubMed  Google Scholar 

  99. Binnington B, Nguyen L, Kamani M, Hossain D, Marks DL, Budani M, et al. Inhibition of Rab prenylation by statins induces cellular glycosphingolipid remodeling. Glycobiology. 2016;26(2):166–80.

    Article  CAS  PubMed  Google Scholar 

  100. Byrne P, Demasi M, Jones M, Smith SM, O’Brien KK, DuBroff R. Evaluating the association between low-density lipoprotein cholesterol reduction and relative and absolute effects of statin treatment: a systematic review and meta-analysis. JAMA Intern Med. 2022;182(5):474–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Avan R, Sahebnasagh A, Hashemi J, Monajati M, Faramarzi F, Henney NC, et al. Update on statin treatment in patients with neuropsychiatric disorders. Life (Basel). 2021;11(12).

  102. Sticozzi C, Belmonte G, Pecorelli A, Cervellati F, Leoncini S, Signorini C, et al. Scavenger receptor B1 post-translational modifications in Rett syndrome. FEBS Lett. 2013;587(14):2199–204.

    Article  CAS  PubMed  Google Scholar 

  103. Tang BL. Cholesterol synthesis inhibition or depletion in axon regeneration. Neural Regen Res. 2022;17(2):271–6.

    Article  CAS  PubMed  Google Scholar 

  104. Asiminas A, Jackson AD, Louros SR, Till SM, Spano T, Dando O, et al. Sustained correction of associative learning deficits after brief, early treatment in a rat model of Fragile X Syndrome. Sci Transl Med. 2019;11(494).

  105. Justice MJ, Buchovecky CM, Kyle SM, Djukic A. A role for metabolism in Rett syndrome pathogenesis: new clinical findings and potential treatment targets. Rare Dis. 2013;1:e27265.

    Article  PubMed  PubMed Central  Google Scholar 

  106. Moazen-Zadeh E, Shirzad F, Karkhaneh-Yousefi MA, Khezri R, Mohammadi MR, Akhondzadeh S. Simvastatin as an adjunctive therapy to risperidone in treatment of autism: a randomized, double-blind, placebo-controlled clinical trial. J Child Adolesc Psychopharmacol. 2018;28(1):82–9.

    Article  CAS  PubMed  Google Scholar 

  107. Stivaros S, Garg S, Tziraki M, Cai Y, Thomas O, Mellor J, et al. Randomised controlled trial of simvastatin treatment for autism in young children with neurofibromatosis type 1 (SANTA). Mol Autism. 2018;9:12.

    Article  PubMed  PubMed Central  Google Scholar 

  108. Dyson A, Ryan M, Garg S, Evans DG, Baines RA. A targeted, low-throughput compound screen in a drosophila model of neurofibromatosis type 1 identifies simvastatin and BMS-204352 as potential therapies for Autism Spectrum Disorder (ASD). eNeuro. 2023;10(5).

  109. Muscas M, Louros SR, Osterweil EK. Lovastatin, not Simvastatin, Corrects Core Phenotypes in the Fragile X Mouse Model. eNeuro. 2019;6(3).

  110. Moffett S, Brown DA, Linder ME. Lipid-dependent targeting of G proteins into rafts. J Biol Chem. 2000;275(3):2191–8.

    Article  CAS  PubMed  Google Scholar 

  111. Brown MS, Goldstein JL. The SREBP pathway: regulation of cholesterol metabolism by proteolysis of a membrane-bound transcription factor. Cell. 1997;89(3):331–40.

    Article  CAS  PubMed  Google Scholar 

  112. Steck TL, Lange Y. SCAP, an ER sensor that regulates cell cholesterol. Dev Cell. 2002;3(3):306–8.

    Article  CAS  PubMed  Google Scholar 

  113. Hosokawa N. Protein degradation assay for endoplasmic reticulum-associated degradation (ERAD) in mammalian cells. In: Nishihara S, Angata K, Aoki-Kinoshita KF, Hirabayashi J, editors. Glycoscience Protocols (GlycoPODv2). Saitama (JP)2021.

  114. Hagiwara M, Nagata K. Redox-dependent protein quality control in the endoplasmic reticulum: folding to degradation. Antioxid Redox Signal. 2012;16(10):1119–28.

    Article  CAS  PubMed  Google Scholar 

  115. El Khouri E, Le Pavec G, Toledano MB, Delaunay-Moisan A. RNF185 is a novel E3 ligase of endoplasmic reticulum-associated degradation (ERAD) that targets cystic fibrosis transmembrane conductance regulator (CFTR). J Biol Chem. 2013;288(43):31177–91.

    Article  PubMed  PubMed Central  Google Scholar 

  116. Pranke IM, Sermet-Gaudelus I. Biosynthesis of cystic fibrosis transmembrane conductance regulator. Int J Biochem Cell Biol. 2014;52:26–38.

    Article  CAS  PubMed  Google Scholar 

  117. Schumacher MM, DeBose-Boyd RA. Posttranslational regulation of HMG CoA reductase, the rate-limiting enzyme in synthesis of cholesterol. Annu Rev Biochem. 2021;90:659–79.

    Article  CAS  PubMed  Google Scholar 

  118. Song BL, Sever N, DeBose-Boyd RA. Gp78, a membrane-anchored ubiquitin ligase, associates with Insig-1 and couples sterol-regulated ubiquitination to degradation of HMG CoA reductase. Mol Cell. 2005;19(6):829–40.

    Article  CAS  PubMed  Google Scholar 

  119. van den Boomen DJH, Volkmar N, Lehner PJ. Ubiquitin-mediated regulation of sterol homeostasis. Curr Opin Cell Biol. 2020;65:103–11.

    Article  PubMed  Google Scholar 

  120. Johnson BM, DeBose-Boyd RA. Underlying mechanisms for sterol-induced ubiquitination and ER-associated degradation of HMG CoA reductase. Semin Cell Dev Biol. 2018;81:121–8.

    Article  CAS  PubMed  Google Scholar 

  121. Tan JME, van der Stoel MM, van den Berg M, van Loon NM, Moeton M, Scholl E, et al. The MARCH6-SQLE axis controls endothelial cholesterol homeostasis and angiogenic sprouting. Cell Rep. 2020;32(5):107944.

    Article  CAS  PubMed  Google Scholar 

  122. Li Y, Gao S, Meng Y. Integrated analysis of endoplasmic reticulum stress regulators’ expression identifies distinct subtypes of autism spectrum disorder. Front Psychiatry. 2023;14:1136154.

    Article  PubMed  PubMed Central  Google Scholar 

  123. He K, Ravindran MS, Tsai B. A bacterial toxin and a nonenveloped virus hijack ER-to-cytosol membrane translocation pathways to cause disease. Crit Rev Biochem Mol Biol. 2015;50(6):477–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Noack J, Bernasconi R, Molinari M. How viruses hijack the ERAD tuning machinery. J Virol. 2014;88(18):10272–5.

    Article  PubMed  PubMed Central  Google Scholar 

  125. Zou L, Wang X, Zhao F, Wu K, Li X, Li Z, et al. Viruses hijack ERAD to regulate their replication and propagation. Int J Mol Sci. 2022;23(16):9398.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Aletrari MO, McKibbin C, Williams H, Pawar V, Pietroni P, Lord JM, et al. Eeyarestatin 1 interferes with both retrograde and anterograde intracellular trafficking pathways. PLoS ONE. 2011;6(7):e22713.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Morito D, Nagata K. Pathogenic Hijacking of ER-Associated Degradation: Is ERAD Flexible? Mol Cell. 2015;59(3):335–44.

    Article  CAS  PubMed  Google Scholar 

  128. Nowakowska-Golacka J, Sominka H, Sowa-Rogozinska N, Slominska-Wojewodzka M. Toxins utilize the endoplasmic reticulum-associated protein degradation pathway in their intoxication process. Int J Mol Sci. 2019;20(6).

  129. Lencer WI, Saslowsky D. Raft trafficking of AB5 subunit bacterial toxins. Biochim Biophys Acta. 2005;1746(3):314–21.

    Article  CAS  PubMed  Google Scholar 

  130. Hazes B, Read RJ. Accumulating evidence suggests that several AB-toxins subvert the endoplasmic reticulum-associated protein degradation pathway to enter target cells. Biochemistry. 1997;36(37):11051–4.

    Article  CAS  PubMed  Google Scholar 

  131. Shi J, Hu X, Guo Y, Wang L, Ji J, Li J, et al. A technique for delineating the unfolding requirements for substrate entry into retrotranslocons during endoplasmic reticulum-associated degradation. J Biol Chem. 2019;294(52):20084–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Dixit G, Mikoryak C, Hayslett T, Bhat A, Draper RK. Cholera toxin up-regulates endoplasmic reticulum proteins that correlate with sensitivity to the toxin. Exp Biol Med (Maywood). 2008;233(2):163–75.

    Article  CAS  PubMed  Google Scholar 

  133. Suzuki Y, Schmitt MJ. Redox diversity in ERAD-mediated protein retrotranslocation from the endoplasmic reticulum: a complex puzzle. Biol Chem. 2015;396(5):539–54.

    Article  CAS  PubMed  Google Scholar 

  134. Burress H, Kellner A, Guyette J, Tatulian SA, Teter K. HSC70 and HSP90 chaperones perform complementary roles in translocation of the cholera toxin A1 subunit from the endoplasmic reticulum to the cytosol. J Biol Chem. 2019;294(32):12122–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Adnan H, Zhang Z, Park HJ, Tailor C, Che C, Kamani M, et al. Endoplasmic reticulum-targeted subunit toxins provide a new approach to rescue misfolded mutant proteins and revert cell models of genetic diseases. PLoS ONE. 2016;11(12):e0166948.

    Article  PubMed  PubMed Central  Google Scholar 

  136. Wang F, Song W, Brancati G, Segatori L. Inhibition of endoplasmic reticulum-associated degradation rescues native folding in loss of function protein misfolding diseases. J Biol Chem. 2011;286(50):43454–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Chen Y, Bellamy WP, Seabra MC, Field MC, Ali BR. ER-associated protein degradation is a common mechanism underpinning numerous monogenic diseases including Robinow syndrome. Hum Mol Genet. 2005;14(17):2559–69.

    Article  CAS  PubMed  Google Scholar 

  138. Douce G, Fontana M, Pizza M, Rappuoli R, Dougan G. Intranasal immunogenicity and adjuvanticity of site-directed mutant derivatives of cholera toxin. Infect Immun. 1997;65(7):2821–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Lingwood C. Therapeutic uses of bacterial subunit toxins. Toxins (Basel). 2021;13(6).

  140. Chen Y, Fan H, Xu C, Hu W, Yu B. Efficient cholera toxin B subunit-based nanoparticles with MRI capability for drug delivery to the brain following intranasal administration. Macromol Biosci. 2019;19(2):e1800340.

    Article  PubMed  Google Scholar 

  141. Guan J, Qian J, Zhan C. Preparation of cholera toxin subunit B functionalized nanoparticles for targeted therapy of glioblastoma. Methods Mol Biol. 2020;2059:207–12.

    Article  CAS  PubMed  Google Scholar 

  142. Schwyzer R. 100 years lock-and-key concept: are peptide keys shaped and guided to their receptors by the target cell membrane? Biopolymers. 1995;37(1):5–16.

    Article  CAS  PubMed  Google Scholar 

  143. Kooistra AJ, de Graaf C, Timmerman H. The receptor concept in 3D: from hypothesis and metaphor to GPCR-ligand structures. Neurochem Res. 2014;39(10):1850–61.

    Article  CAS  PubMed  Google Scholar 

  144. Tusar L, Usenik A, Turk B, Turk D. Mechanisms applied by protein inhibitors to inhibit cysteine proteases. Int J Mol Sci. 2021;22(3).

  145. McGraw KL, Fuhler GM, Johnson JO, Clark JA, Caceres GC, Sokol L, et al. Erythropoietin receptor signaling is membrane raft dependent. PLoS ONE. 2012;7(4):e34477.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Morel E, Ghezzal S, Lucchi G, Truntzer C, Pais de Barros JP, Simon-Plas F, et al. Cholesterol trafficking and raft-like membrane domain composition mediate scavenger receptor class B type 1-dependent lipid sensing in intestinal epithelial cells. Biochim Biophys Acta Mol Cell Biol Lipids. 2018;1863(2):199–211.

  147. Harayama T, Riezman H. Understanding the diversity of membrane lipid composition. Nat Rev Mol Cell Biol. 2018;19(5):281–96.

    Article  CAS  PubMed  Google Scholar 

  148. Edidin M. The state of lipid rafts: from model membranes to cells. Annu Rev Biophys Biomol Struct. 2003;32:257–83.

    Article  CAS  PubMed  Google Scholar 

  149. Yuan C, Johnston LJ. Atomic force microscopy studies of ganglioside GM1 domains in phosphatidylcholine and phosphatidylcholine/cholesterol bilayers. Biophys J. 2001;81(2):1059–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Guo H, Bueler SA, Rubinstein JL. Atomic model for the dimeric F(O) region of mitochondrial ATP synthase. Science. 2017;358(6365):936–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Garcia-Saez AJ, Chiantia S, Schwille P. Effect of line tension on the lateral organization of lipid membranes. J Biol Chem. 2007;282(46):33537–44.

    Article  CAS  PubMed  Google Scholar 

  152. Simons K, Vaz WL. Model systems, lipid rafts, and cell membranes. Annu Rev Biophys Biomol Struct. 2004;33:269–95.

    Article  CAS  PubMed  Google Scholar 

  153. Oradd G, Lindblom G. Lateral diffusion studied by pulsed field gradient NMR on oriented lipid membranes. Magn Reson Chem. 2004;42(2):123–31.

    Article  PubMed  Google Scholar 

  154. Seo S, Murata M, Shinoda W. Pivotal Role of Interdigitation in Interleaflet Interactions: Implications from Molecular Dynamics Simulations. J Phys Chem Lett. 2020;11(13):5171–6.

    Article  CAS  PubMed  Google Scholar 

  155. Suzuki KG, Kasai RS, Hirosawa KM, Nemoto YL, Ishibashi M, Miwa Y, et al. Transient GPI-anchored protein homodimers are units for raft organization and function. Nat Chem Biol. 2012;8(9):774–83.

    Article  CAS  PubMed  Google Scholar 

  156. Sonnino S, Prinetti A, Nakayama H, Yangida M, Ogawa H, Iwabuchi K. Role of very long fatty acid-containing glycosphingolipids in membrane organization and cell signaling: the model of lactosylceramide in neutrophils. Glycoconj J. 2009;26(6):615–21.

    Article  CAS  PubMed  Google Scholar 

  157. Boggs JM, Koshy KM. Do the long fatty acid chains of sphingolipids interdigitate across the center of a bilayer of shorter chain symmetric phospholipids? Biochim Biophys Acta. 1994;1189:233–41.

    Article  CAS  PubMed  Google Scholar 

  158. Ventura AE, Varela ARP, Dingjan T, Santos TCB, Fedorov A, Futerman AH, et al. Lipid domain formation and membrane shaping by C24-ceramide. Biochim Biophys Acta Biomembr. 2020;1862(10): 183400.

    Article  CAS  PubMed  Google Scholar 

  159. Honda A, Nozumi M, Ito Y, Natsume R, Kawasaki A, Nakatsu F, et al. Very-long-chain fatty acids are crucial to neuronal polarity by providing sphingolipids to lipid rafts. Cell Rep. 2023:113195.

  160. Doktorova M, Symons JL, Levental I. Structural and functional consequences of reversible lipid asymmetry in living membranes. Nat Chem Biol. 2020;16(12):1321–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Funding

This manuscript was prepared without funding.

Author information

Authors and Affiliations

Authors

Contributions

C.L. wrote the text and drew the figure.

Corresponding author

Correspondence to Clifford Lingwood.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Competing interests

I am a cofounder of ERAD Therapeutics.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lingwood, C. Is cholesterol both the lock and key to abnormal transmembrane signals in Autism Spectrum Disorder?. Lipids Health Dis 23, 114 (2024). https://doi.org/10.1186/s12944-024-02075-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12944-024-02075-3

Keywords