Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Cell Mol Life Sci. 2009 Jul; 66(14): 2319–2328.
Published online 2009 Apr 16. doi: 10.1007/s00018-009-0022-6
PMCID: PMC2709161
NIHMSID: NIHMS119795
PMID: 19370312

Cytoskeleton–membrane interactions in membrane raft structure

Abstract

Cell membranes are structurally heterogeneous, composed of discrete domains with unique physical and biological properties. Membrane domains can form through a number of mechanisms involving lipid–lipid and protein–lipid interactions. One type of membrane domain is the cholesterol-dependent membrane raft. How rafts form remains a current topic in membrane biology. We review here evidence of structuring of rafts by the cortical actin cytoskeleton. This includes evidence that the actin cytoskeleton associates with rafts, and that many of the structural and functional properties of rafts require an intact actin cytoskeleton. We discuss the mechanisms of the actin-dependent raft organization, and the properties of the actin cytoskeleton in regulating raft-associated signaling events. We end with a discussion of membrane rafts and the actin cytoskeleton in T cell activation, which function synergistically to initiate the adaptive immune response.

Keywords: Membrane rafts, Actin cytoskeleton, Phosphatidylinositol 4, 5 bisphosphate, T cell signaling, Src family kinases

Introduction

Eukaryotic cell membranes consist of a heterogeneous but regulated environment that serves as a dynamic platform for cell functions. Membrane heterogeneity includes discrete membrane domains that form through a combination of protein and lipid interactions [1, 2]. Structurally, membrane domains range from nanoscale structures to molecular assemblies that are microns in size [36]. Research shows that establishing and maintaining this membrane heterogeneity is important for cell viability [710]. Similarly, some pathologies are associated with changes in the structure and composition of membrane domains [1113], and some pathogens utilize membrane domains to gain entry or exit from the target cell [1417]. The functional properties of membrane domains in cell viability and pathology underscore the importance in understanding membrane structure and the mechanisms by which it is established.

One important example of structural and functional membrane domains is the cholesterol-dependent rafts (Fig. 1) [6, 18, 19]. Many of the properties of rafts have been inferred from detergent-resistant membranes (DRMs) that occur in nonionic detergent lysates of animal cells [2, 20, 21]. In the membrane raft model, the DRMs represent poorly solubilized rafts [20], and the composition of the DRMs has served as a guide to the structural and functional properties of rafts. For example, DRMs are enriched with cholesterol and sphingolipids, and it is posited that these lipids provide the structural framework by which the rafts form. Experiments with model membranes show these lipids interact to form a discrete liquid-ordered (Lo) phase in the bilayer [2224]. Furthermore, Lo phase lipids produced in liposomes exhibit many of the physical properties evidenced for rafts in cell membranes, including resistance to solubilization by nonionic detergents and enrichment with proteins and lipids that occur in DRMs [2, 25]. DRM studies have also identified the signals that target proteins to rafts, and these include palmitoylation of a membrane-proximal cysteine, and addition of a glycophosphatidylinositol (GPI) anchor [20, 26].

An external file that holds a picture, illustration, etc.
Object name is 18_2009_22_Fig1_HTML.jpg

Membrane raft model. Cholesterol associations with other membrane lipids, such as sphingolipids, generate a discrete lipid compartment or domain with unique physical and biological properties. The cholesterol confers an ordering on the lipids that imparts changes in the physical properties of the bilayer, including an increase in bilayer width. Through a poorly understood mechanism, the rafts are coupled across the bilayer. Proteins that prefer an ordered lipid environment associate with the domains, often through a discrete targeting signal. A frequent raft-targeting signal for proteins is S-acylation, represented by palmitoylation of a membrane-proximal cysteine. Also enriched in rafts are GPI-anchored proteins. The blue and gray circles represent raft and nonraft lipids, respectively

One shortcoming of the DRM studies is that the detergents are disruptive to membrane structure, and their effects on cell membranes are too complex to draw substantive conclusions regarding the organization of lipids and proteins in the native membrane. For example, detergents used to generate DRMs have been shown to create domains, and to cause mixing and retention of proteins and lipids irrespective of their intrinsic affinity for ordered lipids [27]. Despite the argument that DRMs are detergent artifacts produced during cell lysis [28], in situ imaging studies have provided compelling evidence of nonrandom and clustered distributions for molecules that occur in DRMs [1].

Imaging studies of DRM-associated proteins and lipids show a hierarchical nature to membrane raft structure [1, 3, 4, 29, 30]. The smallest are raft nanoclusters that are approximately 10 nm in diameter, and these are enriched in larger domains that can be several hundred nanometers in size. Even larger are micron-size raft macrodomains that occur in activated cells. The raft macrodomains distinguish themselves as large-scale raft structures by their specific enrichment of proteins and lipids that occur in membrane rafts, and by their resistance to solubilization by nonionic detergents [30, 31]. Examples of raft macrodomains are: the immunological synapse (IS), which forms in activated lymphocytes where they bind to cells bearing antigen [32]; the leading edge and uropod of motile cells [3335]; and cytoskeleton-rich adhesion complexes [36, 37]. These include both cell-to-cell and cell-to-matrix complexes, such as the adherent junctions and focal adhesions [38, 39] [4042].

Structuring of membrane rafts by the actin cytoskeleton

Early descriptions of the membrane raft model evoked the notion that interactions between cholesterol and the sphingolipids generate a lipid platform with which specific proteins associate [2, 21] (Fig. 1). Findings that are more recent, however, suggest that the rafts form in part through capture and stabilization of raft lipids by proteins. As one example, measurements of protein distributions in the plasma membrane showed that clustering of raft-associated proteins was not saturable: the fraction of protein that clustered remained constant with increasing protein expression [43, 44]. This finding suggests that the rafts come about through an ordering of lipids by membrane proteins rather than the rafts occurring as pre-formed lipid complexes with which specific proteins associate. Consistent with this interpretation, protein-dependent ordering of lipids to form Lo phase lipids has been demonstrated in at least two separate studies: cholera toxin B subunit (CTB) binding to liposomes that contain its ligand the ganglioside GM1 [45, 46], and GAP-43 and MARCKS binding to membranes that contain phosphatidylinositol 4, 5 bisphosphate (PIP2) [47].

Refinements of the lipid raft model now ascribe an important role for membrane-associated proteins in forming rafts [43, 48]. One example is the cortical actin cytoskeleton, which is composed of a lattice network of filaments that underlie and attach to the plasma membrane. Proteomic studies show that DRMs are particularly enriched with cytoskeletal proteins, indicative of interactions between the actin cytoskeleton and membrane rafts that could be important in forming and maintaining the rafts. Figure 2 illustrates examples of cytoskeletal proteins that are enriched in rafts, and these include actin, tubulin, myosin, actinin, and supervillin [4954].

An external file that holds a picture, illustration, etc.
Object name is 18_2009_22_Fig2_HTML.jpg

Mechanisms of raft association with the actin cytoskeleton. Measurements of DRMs show rafts are enriched with protein and lipid effectors that function in tethering actin filaments to the plasma membrane. One important example is the lipid cofactor PIP2, which occurs in the cytoplasmic leaflet of the plasma membrane. Another important lipid effector for raft-actin interactions is the phosphoinositide 3-kinase (PI3K)–product PIP3. The PIP3 is necessary for activation of the Rho family GEF Vav. The Rho GTPases activate actin polymerization via Wiskott–Aldrich syndrome family protein (WASP) and WASP family Verprolin-homologous protein (WAVE). These in turn activate actin polymerization and branching through the Arp2/3 complex. Other raft-associated proteins that bind actin filaments are supervillin, myosin-IIA, and myosin IG. Ezrin–radixin–moesin (ERM) proteins link transmembrane proteins, such as adhesion receptors, to the actin cytoskeleton. ERM proteins are regulated by PIP2, which binds the FERM domain of the ERM proteins. Activated integrins associate with membrane rafts. Talin links integrins to the actin cytoskeleton either directly or indirectly by interacting with another cytoskeletal protein vinculin. PIP2 regulates talin interactions with integrins, actin, and vinculin

In a recent study from our laboratory, we compared the role of cholesterol and F-actin in the clustering of membrane-targeted fluorescent proteins by imaging their fluorescence resonance energy transfer (FRET) [29]. First, we observed a co-clustering that was specific to where both the donor and acceptor were associated with rafts. The co-clustering of raft proteins occurred for probes that contained entirely different membrane-targeting signals, thus showing that it was not restricted to one type of raft-targeting signal. As predicted from the membrane raft model, treating the cells with the cholesterol-binding agent filipin specifically inhibited the co-clustering of probes that associate with DRMs. Interestingly, disrupting the actin cytoskeleton with latrunculin B (Lat B) was in some instances more effective than filipin in disrupting co-clustering of raft-associated probes. Finally, co-treating cells with Lat B and filipin resulted in an entirely random distribution of the donor and acceptor probes. Thus, membrane cholesterol and the actin cytoskeleton were sufficient to account for all co-clustering of the raft markers.

Experiments with model membranes show that actin filaments impart an ordering effect on bilayer lipids [55], a property that may underlie structuring of membrane rafts by the actin cytoskeleton. Similarly, environment sensitive lipophilic fluorescent dyes show that the lipids in actin-rich macrodomains of the plasma membrane are condensed in relation to remaining regions of the plasma membrane [56]. This is consistent with the lipids occurring in a relatively ordered state. In Fig. 3 are data that show a decondensation of the plasma membrane in cells treated with Lat B, indicating that an actin-dependent ordering of lipids is a global property of the plasma membrane.

An external file that holds a picture, illustration, etc.
Object name is 18_2009_22_Fig3_HTML.jpg

Global condensation of the plasma membrane by the actin cytoskeleton. The lipophilic dye Laurdan is an environment sensitive fluororophore that serves as a reporter of water penetration into membrane bilayers [126]. Increased lipid packing due to lipid ordering results in a blue shift in the Laurdan emission spectrum, from centered at approximately 500 nm in fluid bilayers to 445 nm in ordered membranes [127]. The normalized ratio of emission channels centered on these wavelengths has been used as a measure of relative membrane ordering [128]. T cells were treated with Laurdan before disrupting the actin cytoskeleton by treating with latrunculin B (Lat B). The samples were imaged in two separate channels, represented by emission wavelengths 400–460 and 470–530 nm using a Leica SP2 multi-photon confocal microscope. General polarization (GP), which reflects the relative lipid condensation [128], is therefore an indicator of lipid ordering. GP was calculated for each pixel in the plasma membrane using the equation GP = I(400–460)−I(470–530)/(I(400–460) + I(470–530)) [56]. GP values range between −1.0 and +1.0, and they are directly proportional to the relative membrane condensation. a Images in the indicated channels of untreated control and Lat B-treated T cells. The bottom row shows the calculated GP image for each sample. b Average GP values of the plasma membrane measured in approximately 50 separate untreated or Lat B-treated T cells

The model in Fig. 4 relates the nano- and micrometer scale clustering of rafts to the structure of the actin cytoskeleton. The nanoclusters are dispersed as complexes associated with filamentous actin. Activation signals bring about an enrichment of actin filaments in the cortical cytoskeleton, which drives the large-scale clustering of rafts to form macrodomains. Also illustrated in Fig. 4 is a network of compartments or corrals established by the actin filaments. This produces a caging effect that can transiently hinder the diffusion of proteins in the plasma membrane. However, the caging is largely nonspecific in nature [5759], and this contrasts with the specific clustering of raft proteins that actin filaments can generate. The contribution of the caging or corralling in establishing rafts is not known, but current evidence suggests that it represents a separate event that occurs on a time scale that is much shorter than the lifetime of the rafts [5860].

An external file that holds a picture, illustration, etc.
Object name is 18_2009_22_Fig4_HTML.jpg

Actin cytoskeleton-dependent raft organization. a Topography of raft nanoclusters and macrodomains. Rafts occur as nanoclusters that are less than 20 nm in diameter, but can cluster to larger structures to include macrodomains that are microns in size. Both the nanoclusters and larger assembles form in an actin-dependent manner [29, 31, 129]. The macrodomains form as a result of signals that activate actin polymerization and attachment of actin filaments to cell membranes. Examples of raft macrodomains include immunological synapses that form in activated lymphocytes, cell–cell adhesion complexes, leading edge and uropod of migrating cells. b The cortical actin cytoskeleton is tethered to the plasma membrane through protein linkers, and these interactions likely structure rafts in the membrane. The meshwork of cortical actin filaments also compartmentalizes the membrane into areas approximately 50–200 nm in size. The compartments represent areas of transient and nonspecific membrane protein confinement by the underlying actin filaments

The notion of an actin-dependent structuring of membrane rafts is an attractive model in that it merges the dynamic properties of the actin cytoskeleton with the important roles that rafts have in membrane functions. Accordingly, signals that cause a localized enrichment and membrane-attachment of actin filaments are predicted to cause a co-enrichment of membrane rafts. This property is exemplified by the raft macrodomains, which form as a result of stimulatory signals that activate actin polymerization and attachment of actin filaments to the plasma membranes [32, 33, 6164]. The macrodomains in turn have specific functions in cell signaling, adhesion, and motility, and the interactions of rafts with the cytoskeleton localize and maintain these functions to discrete regions of the cell surface.

Compartmentalization of PIP2 signaling to membrane rafts

Structuring of rafts by the actin cytoskeleton suggests that proteins and lipids that tether actin filaments to cell membranes are also important for forming rafts. An important regulator of membrane–cytoskeleton interactions is the phosphoinositide PIP2, which serves as a co-factor for many of the proteins that anchor actin filaments to the plasma membrane [65, 66]. Protein binding to PIP2 often occurs through a PIP2-specific recognition sequence, in many cases represented by a PIP2-specific pleckstrin homology (PH) domain [6769]. PIP2-regulated proteins include the ezrin–radixin–moesin (ERM) proteins and talin (Fig. 2) [7073] [74, 75]. The ERM proteins anchor F-actin to membrane proteins that contain a FERM-binding sequence. Talin tethers actin filaments to integrins. Related to PIP2 is the PI3K product PIP3, which also regulates interactions between the cytoskeleton and plasma membrane [76, 77]. For example, PIP3 activates the guanine nucleotide exchange factor (GEF) Vav [78]. Downstream of Vav are the Rho GTPases Rho, Cdc42, and Rac. Cdc42 and Rac activate Wiskott–Aldrich syndrome family protein (WASP) and WASP family Verprolin-homologous protein (WAVE), respectively [7981]. These in turn activate the Arp2/3 complex, which binds F-actin and brings about further actin polymerization and branching of actin filaments [82, 83].

Although PIP2 is enriched in rafts [80, 8486], the notion of raft compartmentalization of its functions to rafts has been controversial. For example, evidence of PIP2 compartmentalization has often been surmised based on an inhibition of its functions by cholesterol-binding agents such as methyl-β-cyclodextrin (MβCD). Drug treatment such as this can produce nonspecific changes in membrane structure that indiscriminately affect membrane functions. Similarly, Jalink and co-workers failed to detect a cholesterol-dependent clustering of PIP2 using FRET to measure the proximity of labeled PIP2-specific PH domains [87].

One recent study from our group detected compartmentalization of PIP2 functions in intact cells by expressing membrane-targeted forms of the PIP2-specific phosphatase Inp54p [88]. Inp54p was targeted to either the raft or nonraft membrane fractions using minimal membrane-anchoring signals, and this selectively increased or decreased the raft pools of PIP2 without changing the global PIP2 content of the cell. Reducing raft PIP2 levels produced cells that had a smooth morphology that was void of the membrane ruffles and filopodia that occurred in normal control cells. Another property was an increase in the number of blebs that formed from the plasma membrane, indicating disassociation of the plasma membrane from the underlying actin cytoskeleton [89]. In contrast, increasing raft PIP2 levels resulted in cells with a striking morphology that included numerous filopodia and extensive membrane ruffling. Enrichment of raft PIP2 also increased cell spreading on poly-l-lysine coated glass surface. Both the cell morphology and cell spreading were sensitive to the PI3K inhibitor wortmannin, suggesting that the phenotype produced by increasing raft PIP2 was due to compartmentalization of either PIP3 or production PIP3-dependent signaling. Interestingly, expression of dominant-negative and constitutively active forms of the Rac generates cell morphologies similar to that which we reported for the raft- and non-raft-targeted Inp54p molecules, respectively [90]. Accordingly, the morphology that was evidenced by increasing raft PIP2 may reflect activation of signals in the Vav–Rac pathway (Fig. 2). This pathway may also give rise to rafts by activating interactions between the cytoskeleton and plasma membrane.

Protein regulation by raft-actin interactions

Membrane rafts in most cell types are enriched with signaling molecules, and a wealth of biochemical and genetic data have provided credence to the notion that rafts function as a specialized signaling platform in cell membranes [26, 9196]. Furthermore, data show that the actin cytoskeleton participates in regulating and activating raft-associated signaling events [9799]. For example, separate studies have shown where protein activity and regulation are tied directly to an intact actin cytoskeleton, and its association with membrane rafts. Examples of this property include G-protein coupled receptors (GPCRs) [100], ERK [101], and Src family kinases (SFKs)[29].

The SFKs illustrate an actin-dependent compartmentalization of protein regulation through sequestering from an important activator. As exemplified by the T cell specific SFK p56lck (Lck), the raft-associated pool of SFK is down-regulated relative to that in the nonraft membrane fraction due to sequestering from its activator CD45 [102]. Additional regulation of SFKs by membrane rafts is achieved through phosphorylation by Csk, which associates with rafts by binding to Csk binding protein (Cbp) or phosphoprotein associated with GEMs (glycolipid-enriched membrane) (PAG) [103]. Experiments from our group showed that treating T cells with Lat B resulted in deregulation of Lck. Furthermore, the changes in Lck regulation were CD45-dependent since the Lat B treatment had no affect on Lck in CD45-deficient T cells [29]. These data together with that regarding actin-dependent clustering of raft proteins suggest a model where membrane-associated actin filaments establish membrane rafts, which then sequester Lck and other raft-associated SFKs from CD45 (Fig. 5). Accordingly, treating T cells with Lat B disrupts the rafts and the sequestering of SFKs from CD45, leading to deregulation of the SFKs. Consistent with this model, filipin treatment also deregulated Lck. The greatest Lck deregulation occurred by co-treating the cells with Lat B and filipin. FRET measurements showed these conditions abolished all co-clustering of raft-associated probes.

An external file that holds a picture, illustration, etc.
Object name is 18_2009_22_Fig5_HTML.jpg

Model for raft protein regulation by cholesterol and the actin cytoskeleton. The Src family kinase (SFK) Lck is activated by the membrane phosphatase CD45 through dephosphorylation of its regulatory C-terminal tyrosine Tyr505. A significant fraction of Lck associates with rafts, yet CD45 is restricted to the nonraft membrane fraction. Raft association of Lck therefore sequesters it from CD45, leading to reduced activity for the raft-associated pool of molecules. Drug treatments that disrupt rafts, such as filipin and Lat B, eliminate this sequestration, resulting in activation of Lck [29]. Conversely, signals that activate actin polymerization and assembly of rafts, such as that from the TCR in T cells, are predicted to enhance the sequestration and down-regulation of raft-associated Lck

Converse to the changes in raft structure and regulation by the drug treatments, activation signals lead to clustering of small raft structures to form large complexes such as the IS [1]. This is predicted to lead to enhance sequestering of CD45 from raft-associated complexes. In the case of T cells, this sequestration is necessary to maintain activation signals since the phosphatase activity of CD45 will quench the TCR-dependent phosphotyrosine signals [104]. Accordingly, T cell activation consists of a choreography where CD45 is first localized at the cell interface with the APC, and then excluded as the IS matures. These properties may underlie interactions between CD45 and SFKs that are first necessary for protein activation, and then removed as the rafts accumulate to form an IS and maintain the activation signals.

Membrane rafts, actin cytoskeleton, and T cell activation

T cell receptor (TCR) signaling exemplifies the interplay between membrane rafts, the actin cytoskeleton, and cell signaling. T cells are activated by antigenic peptide that is presented to the TCR in the context of major histocompatibility complex (MHC) by antigen presenting cells (APCs). TCR signaling is compartmentalized to rafts, and one of the earliest outcomes of these signals is actin polymerization and an actin-driven clustering of rafts [95, 105]. Both the rafts and the signal-driven actin polymerization are necessary for T cell activation; disrupting either the rafts or the actin cytoskeleton inhibits activation [105, 106]. The raft clustering following TCR engagement culminates with formation of an IS where the T cell contacts the APC [107110]. The IS is necessary to sustain the TCR signals [111113]. Disrupting the IS inhibits activation-dependent cell proliferation and cytokine secretion [32].

The co-stimulatory receptor CD28 [114], the integrin LFA-1 [115], and the adhesion receptor CD2 [116] can each provide signals that activate actin polymerization and raft clustering independent of the TCR. In some cases, these signals are sufficient to form an IS independently of the TCR [117]. Furthermore, signals from the TCR alone are often not sufficient to cluster rafts and form an IS. This is particularly the case with naïve T cells, which require a co-stimulatory second signal for their raft clustering [94]. The requirement of TCR and co-stimulatory signals for raft clustering coincides with the two-signal requirement for T cell activation [94, 118]. CD28 is the principal co-stimulatory receptor in naïve T cells, engagement of which is necessary to stimulate IL-2 production [119]. Biochemically, CD28 has no intrinsic activity, but rather functions as a linker protein that recruits specific effectors, including PI3K, Lck, Itk, and filamin A [109, 110, 120, 121]. These effectors also activate actin polymerization or actin-binding to the plasma membrane. However, the hierarchy regarding those most important in providing signals for the actin-dependent raft clustering is largely undefined.

Although the CD28-dependent properties of raft clustering and co-stimulatory signals suggests these events are related, the kinetics by which they occur suggest otherwise. For example, CD28 co-stimulatory signals coincide temporally with those from the TCR [122], which are facile and occur within seconds of its engagement with peptide-bound MHC (pMHC) [123]. However, the actin-driven clustering of rafts takes tens of seconds or minutes [124], indicating that it is too slow to participate proximal to the TCR signaling. Nevertheless, T cell rafts undergo an Ag-independent clustering during initial interactions with the APC as they survey its surface for Ag. We term this early clustering event raft clustering during T cell surveying (RaCS). The RaCS occurs before evidence of TCR-dependent signals, such as TCR stop signals and Ca2+ flux, and it concentrates the CD4 co-receptor and TCR at the APC [125]. Optimization of the T cell membrane environment for TCR signaling by RaCS is predicted to lower the threshold of Ag that is necessary to active the T cell, which is also an important property of co-stimulation. However, it is not known if the early raft clustering event reflects co-stimulatory functions.

Conclusions and perspectives

Recent findings show a synergistic interaction between the actin cytoskeleton and membrane rafts. Actin filaments closely associate with the plasma membrane in many cellular processes. Membrane rafts are enriched with the modulators of cortical actin cytoskeleton and with factors that anchor actin filaments to the plasma membrane. Furthermore, the actin cytoskeleton regulates the clustering of membrane raft proteins in a specific manner and at nanoscale level of membrane structure. The actin cytoskeleton is a highly dynamic structure, and its role in establishing membrane rafts provides one avenue for tuning the membrane microenvironment in a manner that favors or inhibits discrete membrane functions. However, further studies are necessary to better resolve the mechanism for the actin-dependent structuring of membrane rafts.

Acknowledgments

This work was supported by NIH grant R01 GM070001 and Oklahoma Center for the advancement of Science and Technology grants HR08-084 (WR).

References

1. Rodgers W, Farris D, Mishra S. Merging complexes: properties of membrane raft assembly during lymphocyte signaling. Trends Immunol. 2005;26:97–103. doi: 10.1016/j.it.2004.11.016. [PubMed] [CrossRef] [Google Scholar]
2. Brown DA, London E. Structure and function of sphingolipid- and cholesterol-rich membrane rafts. J Biol Chem. 2000;275:17221–17224. doi: 10.1074/jbc.R000005200. [PubMed] [CrossRef] [Google Scholar]
3. Sharma P, Varma R, Sarasij RC, Ira, Gousset K, Krishnamoorthy G, Rao M, Mayor S (2004) Nanoscale organization of multiple GPI-anchored proteins in living cell membranes. Cell 116: 577–589 [PubMed]
4. Zacharias DA, Violin JD, Newton AC, Tsien RY. Partitioning of lipid-modified monomeric GFPs into membrane microdomains of live cells. Science. 2002;296:913–916. doi: 10.1126/science.1068539. [PubMed] [CrossRef] [Google Scholar]
5. Eggeling C, Ringemann C, Medda R, Schwarzmann G, Sandhoff K, Polyakova S, Belov VN, Hein B, von Middendorff C, Schonle A, Hell SW. Direct observation of the nanoscale dynamics of membrane lipids in a living cell. Nature. 2009;457:1159–1162. doi: 10.1038/nature07596. [PubMed] [CrossRef] [Google Scholar]
6. Jacobson K, Mouritsen OG, Anderson RG. Lipid rafts: at a crossroad between cell biology and physics. Nat Cell Biol. 2007;9:7–14. doi: 10.1038/ncb0107-7. [PubMed] [CrossRef] [Google Scholar]
7. Yang B, Oo TN, Rizzo V. Lipid rafts mediate H2O2 prosurvival effects in cultured endothelial cells. FASEB J. 2006;20:1501–1503. doi: 10.1096/fj.05-5359fje. [PubMed] [CrossRef] [Google Scholar]
8. Yamazaki S, Iwama A, Takayanagi S, Morita Y, Eto K, Ema H, Nakauchi H. Cytokine signals modulated via lipid rafts mimic niche signals and induce hibernation in hematopoietic stem cells. EMBO J. 2006;25:3515–3523. doi: 10.1038/sj.emboj.7601236. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
9. Furne C, Corset V, Herincs Z, Cahuzac N, Hueber AO, Mehlen P. The dependence receptor DCC requires lipid raft localization for cell death signaling. Proc Natl Acad Sci USA. 2006;103:4128–4133. doi: 10.1073/pnas.0507864103. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
10. Koenig A, Russell JQ, Rodgers WA, Budd RC. Spatial differences in active caspase-8 defines its role in T cell activation versus cell death. Cell Death Differ. 2008;15:1701–1711. doi: 10.1038/cdd.2008.100. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
11. Baruthio F, Quadroni M, Ruegg C, Mariotti A. Proteomic analysis of membrane rafts of melanoma cells identifies protein patterns characteristic of the tumor progression stage. Proteomics. 2008;8:4733–4747. doi: 10.1002/pmic.200800169. [PubMed] [CrossRef] [Google Scholar]
12. Jury EC, Kabouridis PS, Flores-Borja F, Mageed RA, Isenberg DA. Altered lipid raft-associated signaling and ganglioside expression in T lymphocytes from patients with systemic lupus erythematosus. J Clin Invest. 2004;113:1176–1187. [PMC free article] [PubMed] [Google Scholar]
13. Banfi C, Brioschi M, Wait R, Begum S, Gianazza E, Fratto P, Polvani G, Vitali E, Parolari A, Mussoni L, Tremoli E. Proteomic analysis of membrane microdomains derived from both failing and non-failing human hearts. Proteomics. 2006;6:1976–1988. doi: 10.1002/pmic.200500278. [PubMed] [CrossRef] [Google Scholar]
14. Manes S, del Real G, Martinez AC. Pathogens: raft hijackers. Nat Rev Immunol. 2003;3:557–568. doi: 10.1038/nri1129. [PubMed] [CrossRef] [Google Scholar]
15. Marsh M, Helenius A. Virus entry: open sesame. Cell. 2006;124:729–740. doi: 10.1016/j.cell.2006.02.007. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
16. Cambi A, de Lange F, van Maarseveen NM, Nijhuis M, Joosten B, van Dijk EM, de Bakker BI, Fransen JA, Bovee-Geurts PH, van Leeuwen FN, Van Hulst NF, Figdor CG. Microdomains of the C-type lectin DC-SIGN are portals for virus entry into dendritic cells. J Cell Biol. 2004;164:145–155. doi: 10.1083/jcb.200306112. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
17. Bhattacharya B, Roy P. Bluetongue virus outer capsid protein VP5 interacts with membrane lipid rafts via a snare domain. J Virol. 2008;82:10600–10612. doi: 10.1128/JVI.01274-08. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
18. Simons K, Ikonen E. Functional rafts in cell membranes. Nature. 1997;387:569–572. doi: 10.1038/42408. [PubMed] [CrossRef] [Google Scholar]
19. Sengupta P, Baird B, Holowka D. Lipid rafts, fluid/fluid phase separation, and their relevance to plasma membrane structure and function. Semin Cell Dev Biol. 2007;18:583–590. doi: 10.1016/j.semcdb.2007.07.010. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
20. Brown DA, Rose JK. Sorting of GPI-anchored proteins to glycolipid-enriched membrane subdomains during transport to the apical cell surface. Cell. 1992;68:533–544. doi: 10.1016/0092-8674(92)90189-J. [PubMed] [CrossRef] [Google Scholar]
21. Simons K, Toomre D. Lipid rafts and signal transduction. Nat Mol Cell Biol Rev. 2000;1:31–39. doi: 10.1038/35036052. [PubMed] [CrossRef] [Google Scholar]
22. Chiantia S, Kahya N, Schwille P. Raft domain reorganization driven by short- and long-chain ceramide: a combined AFM and FCS study. Langmuir. 2007;23:7659–7665. doi: 10.1021/la7010919. [PubMed] [CrossRef] [Google Scholar]
23. Risselada HJ, Marrink SJ. The molecular face of lipid rafts in model membranes. Proc Natl Acad Sci USA. 2008;105:17367–17372. doi: 10.1073/pnas.0807527105. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
24. Sankaram MB, Thompson TE. Interaction of cholesterol with various glycerophospholipids and sphingomyelin. Biochemistry. 1990;29:10670–10675. doi: 10.1021/bi00499a014. [PubMed] [CrossRef] [Google Scholar]
25. Schroeder R, London E, Brown D. Interactions between saturated acyl chains confer detergent resistance on lipids and glycosylphosphatidylinositol (GPI)-anchored proteins: GPI-anchored proteins in liposomes and cells show similar behavior. Proc Natl Acad Sci USA. 1994;91:12130–12134. doi: 10.1073/pnas.91.25.12130. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
26. Rodgers W, Crise B, Rose JK. Signals determining protein tyrosine kinase and glycosyl-phosphatidylinositol-anchored protein targeting to a glycolipid-enriched membrane fraction. Mol Cell Biol. 1994;14:5384–5391. [PMC free article] [PubMed] [Google Scholar]
27. Heerklotz H. Triton promotes domain formation in lipid raft mixtures. Biophys J. 2002;83:2693–2701. doi: 10.1016/S0006-3495(02)75278-8. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
28. Munro S. Lipid rafts: elusive or illusive? Cell. 2003;115:377–388. doi: 10.1016/S0092-8674(03)00882-1. [PubMed] [CrossRef] [Google Scholar]
29. Chichili GR, Rodgers W. Clustering of membrane raft proteins by the actin cytoskeleton. J Biol Chem. 2007;282:36682–36691. doi: 10.1074/jbc.M702959200. [PubMed] [CrossRef] [Google Scholar]
30. Jordan S, Rodgers W. T cell glycolipid-enriched membrane domains are constitutively assembled as membrane patches that translocate to immune synapses. J Immunol. 2003;171:78–87. [PubMed] [Google Scholar]
31. Rodgers W, Zavzavadjian J. Glycolipid-enriched membrane domains are assembled into membrane patches by associating with the actin cytoskeleton. Exp Cell Res. 2001;267:173–183. doi: 10.1006/excr.2001.5253. [PubMed] [CrossRef] [Google Scholar]
32. Grakoui A, Bromley SK, Sumen C, Davis MM, Shaw AS, Allen PM, Dustin ML. The immunological synapse: a molecular machine controlling T cell activation. Science. 1999;285:221–227. doi: 10.1126/science.285.5425.221. [PubMed] [CrossRef] [Google Scholar]
33. Golub T, Caroni P. Pi(4, 5)P2-dependent microdomain assemblies capture microtubules to promote and control leading edge motility. J Cell Biol. 2005;169:151–165. doi: 10.1083/jcb.200407058. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
34. Gomez-Mouton C, Abad JL, Mira E, Lacalle RA, Gallardo E, Jimenez-Baranda S, Illa I, Bernad A, Manes S, Martinez AC. Segregation of leading-edge and uropod components into specific lipid rafts during t cell polarization. Proc Natl Acad Sci USA. 2001;98:9642–9647. doi: 10.1073/pnas.171160298. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
35. Gomez-Mouton C, Lacalle RA, Mira E, Jimenez-Baranda S, Barber DF, Carrera AC, Martinez AC, Manes S. Dynamic redistribution of raft domains as an organizing platform for signaling during cell chemotaxis. J Cell Biol. 2004;164:759–768. doi: 10.1083/jcb.200309101. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
36. Gaus K, Le Lay S, Balasubramanian N, Schwartz MA. Integrin-mediated adhesion regulates membrane order. J Cell Biol. 2006;174:725–734. doi: 10.1083/jcb.200603034. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
37. Krauss K, Altevogt P. Integrin leukocyte function-associated antigen-1-mediated cell binding can be activated by clustering of membrane rafts. J Biol Chem. 1999;274:36921–36927. doi: 10.1074/jbc.274.52.36921. [PubMed] [CrossRef] [Google Scholar]
38. Causeret M, Taulet N, Comunale F, Favard C, Gauthier-Rouviere C. N-cadherin association with lipid rafts regulates its dynamic assembly at cell–cell junctions in C2C12 myoblasts. Mol Biol Cell. 2005;16:2168–2180. doi: 10.1091/mbc.E04-09-0829. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
39. Lambert M, Thoumine O, Brevier J, Choquet D, Riveline D, Mege RM. Nucleation and growth of cadherin adhesions. Exp Cell Res. 2007;313:4025–4040. doi: 10.1016/j.yexcr.2007.07.035. [PubMed] [CrossRef] [Google Scholar]
40. Decker L, ffrench-Constant C. Lipid rafts and integrin activation regulate oligodendrocyte survival. J Neurosci. 2004;24:3816–3825. doi: 10.1523/JNEUROSCI.5725-03.2004. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
41. Porter JC, Hogg N. Integrins take partners: cross-talk between integrins and other membrane receptors. Trends Cell Biol. 1998;8:390–396. doi: 10.1016/S0962-8924(98)01344-0. [PubMed] [CrossRef] [Google Scholar]
42. Lehembre F, Yilmaz M, Wicki A, Schomber T, Strittmatter K, Ziegler D, Kren A, Went P, Derksen PW, Berns A, Jonkers J, Christofori G. NCAM-induced focal adhesion assembly: a functional switch upon loss of E-cadherin. EMBO J. 2008;27:2603–2615. doi: 10.1038/emboj.2008.178. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
43. Plowman SJ, Muncke C, Parton RG, Hancock JF. H-ras, K-ras, and inner plasma membrane raft proteins operate in nanoclusters with differential dependence on the actin cytoskeleton. Proc Natl Acad Sci USA. 2005;102:15500–15505. doi: 10.1073/pnas.0504114102. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
44. Tian T, Harding A, Inder K, Plowman S, Parton RG, Hancock JF. Plasma membrane nanoswitches generate high-fidelity ras signal transduction. Nat Cell Biol. 2007;9:905–914. doi: 10.1038/ncb1615. [PubMed] [CrossRef] [Google Scholar]
45. Hammond AT, Heberle FA, Baumgart T, Holowka D, Baird B, Feigenson GW. Crosslinking a lipid raft component triggers liquid ordered-liquid disordered phase separation in model plasma membranes. Proc Natl Acad Sci USA. 2005;102:6320–6325. doi: 10.1073/pnas.0405654102. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
46. Forstner MB, Yee CK, Parikh AN, Groves JT. Lipid lateral mobility and membrane phase structure modulation by protein binding. J Am Chem Soc. 2006;128:15221–15227. doi: 10.1021/ja064093h. [PubMed] [CrossRef] [Google Scholar]
47. Tong J, Nguyen L, Vidal A, Simon SA, Skene JH, McIntosh TJ. Role of GAP-43 in sequestering phosphatidylinositol 4, 5-bisphosphate to raft bilayers. Biophys J. 2008;94:125–133. doi: 10.1529/biophysj.107.110536. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
48. Douglass AD, Vale RD. Single-molecule microscopy reveals plasma membrane microdomains created by protein–protein networks that exclude or trap signaling molecules in T cells. Cell. 2005;121:937–950. doi: 10.1016/j.cell.2005.04.009. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
49. Nebl T, Pestonjamasp KN, Leszyk JD, Crowley JL, Oh SW, Luna EJ. Proteomic analysis of a detergent-resistant membrane skeleton from neutrophil plasma membranes. J Biol Chem. 2002;277:43399–43409. doi: 10.1074/jbc.M205386200. [PubMed] [CrossRef] [Google Scholar]
50. Bini L, Pacini S, Liberatori S, Valensin S, Pellegrini M, Raggiaschi R, Pallini V, Baldari CT. Extensive temporally regulated reorganization of the lipid raft proteome following T-cell antigen receptor triggering. Biochem J. 2003;369:301–309. doi: 10.1042/BJ20020503. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
51. von Haller PD, Donohoe S, Goodlett DR, Aebersold R, Watts JD. Mass spectrometric characterization of proteins extracted from jurkat T cell detergent-resistant membrane domains. Proteomics. 2001;1:1010–1021. doi: 10.1002/1615-9861(200108)1:8<1010::AID-PROT1010>3.0.CO;2-L. [PubMed] [CrossRef] [Google Scholar]
52. Yanagida M, Nakayama H, Yoshizaki F, Fujimura T, Takamori K, Ogawa H, Iwabuchi K. Proteomic analysis of plasma membrane lipid rafts of HL-60 cells. Proteomics. 2007;7:2398–2409. doi: 10.1002/pmic.200700056. [PubMed] [CrossRef] [Google Scholar]
53. Yu MJ, Pisitkun T, Wang G, Aranda JF, Gonzales PA, Tchapyjnikov D, Shen RF, Alonso MA, Knepper MA. Large-scale quantitative LC-MS/MS analysis of detergent-resistant membrane proteins from rat renal collecting duct. Am J Physiol Cell Physiol. 2008;295:C661–C678. doi: 10.1152/ajpcell.90650.2007. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
54. MacLellan DL, Steen H, Adam RM, Garlick M, Zurakowski D, Gygi SP, Freeman MR, Solomon KR. A quantitative proteomic analysis of growth factor-induced compositional changes in lipid rafts of human smooth muscle cells. Proteomics. 2005;5:4733–4742. doi: 10.1002/pmic.200500044. [PubMed] [CrossRef] [Google Scholar]
55. Liu AP, Fletcher DA. Actin polymerization serves as a membrane domain switch in model lipid bilayers. Biophys J. 2006;91:4064–4070. doi: 10.1529/biophysj.106.090852. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
56. Gaus K, Chklovskaia E, Fazekas de St Groth B, Jessup W, Harder T. Condensation of the plasma membrane at the site of T lymphocyte activation. J Cell Biol. 2005;171:121–131. doi: 10.1083/jcb.200505047. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
57. Kusumi A, Nakada C, Ritchie K, Murase K, Suzuki K, Murakoshi H, Kasai RS, Kondo J, Fujiwara T. Paradigm shift of the plasma membrane concept from the two-dimensional continuum fluid to the partitioned fluid: high-speed single-molecule tracking of membrane molecules. Annu Rev Biophys Biomol Struct. 2005;34:351–378. doi: 10.1146/annurev.biophys.34.040204.144637. [PubMed] [CrossRef] [Google Scholar]
58. Suzuki KG, Fujiwara TK, Edidin M, Kusumi A. Dynamic recruitment of phospholipase C γ at transiently immobilized GPI-anchored receptor clusters induces IP3-Ca2+ signaling: single-molecule tracking study 2. J Cell Biol. 2007;177:731–742. doi: 10.1083/jcb.200609175. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
59. Suzuki KG, Fujiwara TK, Sanematsu F, Iino R, Edidin M, Kusumi A. GPI-anchored receptor clusters transiently recruit Lyn and Gα for temporary cluster immobilization and Lyn activation: single-molecule tracking study 1. J Cell Biol. 2007;177:717–730. doi: 10.1083/jcb.200609174. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
60. Hess ST, Gould TJ, Gudheti MV, Maas SA, Mills KD, Zimmerberg J. Dynamic clustered distribution of hemagglutinin resolved at 40 nm in living cell membranes discriminates between raft theories. Proc Natl Acad Sci USA. 2007;104:17370–17375. doi: 10.1073/pnas.0708066104. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
61. Roumier A, Olivo-Marin JC, Arpin M, Michel F, Martin M, Mangeat P, Acuto O, Dautry-Varsat A, Alcover A. The membrane-microfilament linker ezrin is involved in the formation of the immunological synapse and in T cell activation. Immunity. 2001;15:715–728. doi: 10.1016/S1074-7613(01)00225-4. [PubMed] [CrossRef] [Google Scholar]
62. Allenspach EJ, Cullinan P, Tong J, Tang Q, Tesciuba AG, Cannon JL, Takahashi SM, Morgan R, Burkhardt JK, Sperling AI. ERM-dependent movement of CD43 defines a novel protein complex distal to the immunological synapse. Immunity. 2001;15:739–750. doi: 10.1016/S1074-7613(01)00224-2. [PubMed] [CrossRef] [Google Scholar]
63. Lee JL, Wang MJ, Sudhir PR, Chen JY. CD44 engagement promotes matrix-derived survival through the CD44-Src-integrin axis in lipid rafts. Mol Cell Biol. 2008;28:5710–5723. doi: 10.1128/MCB.00186-08. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
64. Balasubramanian N, Scott DW, Castle JD, Casanova JE, Schwartz MA. Arf6 and microtubules in adhesion-dependent trafficking of lipid rafts. Nat Cell Biol. 2007;9:1381–1391. doi: 10.1038/ncb1657. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
65. Czech MP. PIP2 and PIP3: Complex roles at the cell surface. Cell. 2000;100:603–606. doi: 10.1016/S0092-8674(00)80696-0. [PubMed] [CrossRef] [Google Scholar]
66. Yin HL, Janmey PA. Phosphoinositide regulation of the actin cytoskeleton. Annu Rev Physiol. 2003;65:761–789. doi: 10.1146/annurev.physiol.65.092101.142517. [PubMed] [CrossRef] [Google Scholar]
67. Varnai P, Balla T. Visualization of phosphoinositides that bind pleckstrin homology domains: Calcium- and agonist-induced dynamic changes and relationship to myo-[3H]inositol-labeled phosphoinositide pools. J Cell Biol. 1998;143:501–510. doi: 10.1083/jcb.143.2.501. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
68. Harlan JE, Hajduk PJ, Yoon HS, Fesik SW. Pleckstrin homology domains bind to phosphatidylinositol-4, 5-bisphosphate. Nature. 1994;371:168–170. doi: 10.1038/371168a0. [PubMed] [CrossRef] [Google Scholar]
69. Lemon G, Gibson WG, Bennett MR. Metabotropic receptor activation, desensitization and sequestration-II: modelling the dynamics of the pleckstrin homology domain. J Theor Biol. 2003;223:113–129. doi: 10.1016/S0022-5193(03)00080-8. [PubMed] [CrossRef] [Google Scholar]
70. Hirao M, Sato N, Kondo T, Yonemura S, Monden M, Sasaki T, Takai Y, Tsukita S, Tsukita S. Regulation mechanism of ERM (ezrin/radixin/moesin) protein/plasma membrane association: possible involvement of phosphatidylinositol turnover and Rho-dependent signaling pathway. J Cell Biol. 1996;135:37–51. doi: 10.1083/jcb.135.1.37. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
71. Heiska L, Alfthan K, Gronholm M, Vilja P, Vaheri A, Carpen O. Association of ezrin with intercellular adhesion molecule-1 and -2 (ICAM-1 and ICAM-2). Regulation by phosphatidylinositol 4, 5-bisphosphate. J Biol Chem. 1998;273:21893–21900. doi: 10.1074/jbc.273.34.21893. [PubMed] [CrossRef] [Google Scholar]
72. Blin G, Margeat E, Carvalho K, Royer CA, Roy C, Picart C. Quantitative analysis of the binding of ezrin to large unilamellar vesicles containing phosphatidylinositol 4, 5 bisphosphate. Biophys J. 2008;94:1021–1033. doi: 10.1529/biophysj.107.110213. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
73. Zhang M, Bohlson SS, Dy M, Tenner AJ. Modulated interaction of the ERM protein, moesin, with CD93. Immunology. 2005;115:63–73. doi: 10.1111/j.1365-2567.2005.02120.x. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
74. Kaufmann S, Kas J, Goldmann WH, Sackmann E, Isenberg G. Talin anchors and nucleates actin filaments at lipid membranes. A direct demonstration. FEBS Lett. 1992;314:203–205. doi: 10.1016/0014-5793(92)80975-M. [PubMed] [CrossRef] [Google Scholar]
75. Martel V, Racaud-Sultan C, Dupe S, Marie C, Paulhe F, Galmiche A, Block MR, Albiges-Rizo C. Conformation, localization, and integrin binding of talin depend on its interaction with phosphoinositides. J Biol Chem. 2001;276:21217–21227. doi: 10.1074/jbc.M102373200. [PubMed] [CrossRef] [Google Scholar]
76. Villalba M, Bi K, Rodriguez F, Tanaka Y, Schoenberger S, Altman A. Vav1/Rac-dependent actin cytoskeleton reorganization is required for lipid raft clustering in T cells. J Cell Biol. 2001;155:331–338. doi: 10.1083/jcb.200107080. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
77. Inabe K, Ishiai M, Scharenberg AM, Freshney N, Downward J, Kurosaki T. Vav3 modulates B cell receptor responses by regulating phosphoinositide 3-kinase activation. J Exp Med. 2002;195:189–200. doi: 10.1084/jem.20011571. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
78. Bustelo XR. Regulatory and signaling properties of the Vav family. Mol Cell Biol. 2000;20:1461–1477. doi: 10.1128/MCB.20.5.1461-1477.2000. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
79. Higgs HN, Pollard TD. Activation by Cdc42 and PiP2 of Wiskott–Aldrich syndrome protein (WASp) stimulates actin nucleation by Arp2/3 complex. J Cell Biol. 2000;150:1311–1320. doi: 10.1083/jcb.150.6.1311. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
80. Rozelle AL, Machesky LM, Yamamoto M, Driessens MH, Insall RH, Roth MG, Luby-Phelps K, Marriott G, Hall A, Yin HL. Phosphatidylinositol 4, 5-bisphosphate induces actin-based movement of raft-enriched vesicles through WASp–Arp2/3. Curr Biol. 2000;10:311–320. doi: 10.1016/S0960-9822(00)00384-5. [PubMed] [CrossRef] [Google Scholar]
81. Lai FP, Szczodrak M, Block J, Faix J, Breitsprecher D, Mannherz HG, Stradal TE, Dunn GA, Small JV, Rottner K. Arp2/3 complex interactions and actin network turnover in lamellipodia. EMBO J. 2008;27:982–992. doi: 10.1038/emboj.2008.34. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
82. Goley ED, Welch MD. The Arp2/3 complex: an actin nucleator comes of age. Nat Rev Mol Cell Biol. 2006;7:713–726. doi: 10.1038/nrm2026. [PubMed] [CrossRef] [Google Scholar]
83. Tseng Y, Wirtz D. Dendritic branching and homogenization of actin networks mediated by Arp2/3 complex. Phys Rev Lett. 2004;93:258104. doi: 10.1103/PhysRevLett.93.258104. [PubMed] [CrossRef] [Google Scholar]
84. Hope HR, Pike LJ. Phosphoinositides and phosphoinositide-utilizing enzymes in detergent-insoluble lipid domains. Mol Biol Cell. 1996;7:843–851. [PMC free article] [PubMed] [Google Scholar]
85. Laux T, Fukami K, Thelen M, Golub T, Frey D, Caroni P. GAP43, MARCKS, and CAP23 modulate PI(4, 5)P2 at plasmalemmal rafts, and regulate cell cortex actin dynamics through a common mechanism. J Cell Biol. 2000;149:1455–1472. doi: 10.1083/jcb.149.7.1455. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
86. Parmryd I, Adler J, Patel R, Magee AI. Imaging metabolism of phosphatidylinositol 4, 5-bisphosphate in T-cell GM1-enriched domains containing ras proteins. Exp Cell Res. 2003;285:27–38. doi: 10.1016/S0014-4827(02)00048-4. [PubMed] [CrossRef] [Google Scholar]
87. van Rheenen J, Achame EM, Janssen H, Calafat J, Jalink K. PIP2 signaling in lipid domains: a critical re-evaluation. EMBO J. 2005;24:1664–1673. doi: 10.1038/sj.emboj.7600655. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
88. Johnson CM, Chichili GR, Rodgers W. Compartmentalization of phosphatidylinositol 4, 5-bisphosphate signaling evidenced using targeted phosphatases. J Biol Chem. 2008;283:29920–29928. doi: 10.1074/jbc.M805921200. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
89. Charras CT, Hu CK, Coughlin M, Mitchison TJ. Reassembly of contractile actin cortex in cell blebs. J Cell Biol. 2006;175:477–490. doi: 10.1083/jcb.200602085. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
90. Arrieumerlou C, Randriamampita C, Bismuth G, Trautmann A. Rac is involved in early TCR signaling. J. Immunol. 2000;165:3182–3189. [PubMed] [Google Scholar]
91. Zhang W, Trible RP, Samelson LE. LAT palmitoylation: its essential role in membrane microdomain targeting and tyrosine phosphorylation during T cell activation. Immunity. 1998;9:239–246. doi: 10.1016/S1074-7613(00)80606-8. [PubMed] [CrossRef] [Google Scholar]
92. Brdicka T, Pavlistova D, Leo A, Bruyns E, Korinek V, Angelisova P, Scherer J, Shevchenko A, Hilgert I, Cerny J, Drbal K, Kuramitsu Y, Kornacker B, Horejsi V, Schraven B. Phosphoprotein associated with glycosphingolipid-enriched microdomains (PAG), a novel ubiquitously expressed transmembrane adaptor protein, binds the protein tyrosine kinase Csk and is involved in regulation of T cell activation. J Exp Med. 2000;191:1591–1604. doi: 10.1084/jem.191.9.1591. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
93. Roy S, Luetterforst R, Harding A, Apolloni A, Etheridge M, Stang E, Rolls B, Hancock JF, Parton RG. Dominant-negative caveolin inhibits H-ras function by disrupting cholesterol-rich plasma membrane domains. Nat Cell Biol. 1999;1:98–105. doi: 10.1038/15687. [PubMed] [CrossRef] [Google Scholar]
94. Viola A, Schroeder S, Sakakibara Y, Lanzavecchia A. T lymphocyte costimulation mediated by reorganization of membrane microdomains. Science. 1999;283:680–682. doi: 10.1126/science.283.5402.680. [PubMed] [CrossRef] [Google Scholar]
95. Xavier R, Brennan T, Li Q, McCormack C, Seed B. Membrane compartmentation is required for efficient T cell activation. Immunity. 1998;8:723–732. doi: 10.1016/S1074-7613(00)80577-4. [PubMed] [CrossRef] [Google Scholar]
96. Kabouridis PS, Magee AI, Ley SC. S-acylation of Lck protein tyrosine kinase is essential for its signalling function in T lymphocytes. EMBO J. 1997;16:4983–4998. doi: 10.1093/emboj/16.16.4983. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
97. Bunnell SC, Kapoor V, Trible RP, Zhang W, Samelson LE. Dynamic actin polymerization drives T cell receptor-induced spreading: A role for the signal transduction adaptor LAT. Immunity. 2001;14:315–329. doi: 10.1016/S1074-7613(01)00112-1. [PubMed] [CrossRef] [Google Scholar]
98. Barda-Saad M, Braiman A, Titerence R, Bunnell SC, Barr VA, Samelson LE. Dynamic molecular interactions linking the T cell antigen receptor to the actin cytoskeleton. Nat Immunol. 2005;6:80–89. doi: 10.1038/ni1143. [PubMed] [CrossRef] [Google Scholar]
99. Gomez TS, Billadeau DD. T cell activation and the cytoskeleton: you can’t have one without the other. Adv Immunol. 2008;97:1–64. doi: 10.1016/S0065-2776(08)00001-1. [PubMed] [CrossRef] [Google Scholar]
100. Sah VP, Seasholtz TM, Sagi SA, Brown JH. The role of Rho in G protein-coupled receptor signal transduction. Annu Rev Pharmacol Toxicol. 2000;40:459–489. doi: 10.1146/annurev.pharmtox.40.1.459. [PubMed] [CrossRef] [Google Scholar]
101. Pullikuth AK, Catling AD. Scaffold mediated regulation of MAPK signaling and cytoskeletal dynamics: A perspective. Cell Signal. 2007;19:1621–1632. doi: 10.1016/j.cellsig.2007.04.012. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
102. Rodgers W, Rose JK. Exclusion of CD45 inhibits activity of p56Lck associated with glycolipid-enriched membrane domains. J Cell Biol. 1996;135:1515–1523. doi: 10.1083/jcb.135.6.1515. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
103. Davidson D, Bakinowski M, Thomas ML, Horejsi V, Veillette A. Phosphorylation-dependent regulation of T-cell activation by PAG/Cbp, a lipid raft-associated transmembrane adaptor. Mol Cell Biol. 2003;23:2017–2028. doi: 10.1128/MCB.23.6.2017-2028.2003. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
104. Freiberg BA, Kupfer H, Maslanik W, Delli J, Kappler J, Zaller DM, Kupfer A. Staging and resetting T cell activation in SMACs. Nat Immunol. 2002;3:911–917. doi: 10.1038/ni836. [PubMed] [CrossRef] [Google Scholar]
105. Miceli MC, Moran M, Chung CD, Patel VP, Low T, Zinnanti W. Co-stimulation and counter-stimulation: lipid raft clustering controls TCR signaling and functional outcomes. Semin Immunol. 2001;13:115–128. doi: 10.1006/smim.2000.0303. [PubMed] [CrossRef] [Google Scholar]
106. Wulfing C, Purtic B, Klem J, Schatzle JD. Stepwise cytoskeletal polarization as a series of checkpoints in innate but not adaptive cytolytic killing. Proc Natl Acad Sci USA. 2003;100:7767–7772. doi: 10.1073/pnas.1336920100. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
107. Montixi C, Langlet C, Bernard AM, Thimonier J, Dubois C, Wurbel MA, Chauvin JP, Pierres M, He HT. Engagement of T cell receptor triggers its recruitment to low-density detergent-insoluble membrane domains. EMBO J. 1998;17:5334–5348. doi: 10.1093/emboj/17.18.5334. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
108. Janes PW, Ley SC, Magee AI. Aggregation of lipid rafts accompanies signaling via the T cell antigen receptor. J Cell Biol. 1999;147:447–461. doi: 10.1083/jcb.147.2.447. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
109. Tavano R, Contento RL, Baranda SJ, Soligo M, Tuosto L, Manes S, Viola A. CD28 interaction with filamin-A controls lipid raft accumulation at the T-cell immunological synapse. Nat Cell Biol. 2006;8:1270–1276. doi: 10.1038/ncb1492. [PubMed] [CrossRef] [Google Scholar]
110. Tavano R, Gri G, Molon B, Marinari B, Rudd CE, Tuosto L, Viola A. CD28 and lipid rafts coordinate recruitment of Lck to the immunological synapse of human T lymphocytes. J Immunol. 2004;173:5392–5397. [PubMed] [Google Scholar]
111. Cemerski S, Das J, Giurisato E, Markiewicz MA, Allen PM, Chakraborty AK, Shaw AS. The balance between T cell receptor signaling and degradation at the center of the immunological synapse is determined by antigen quality. Immunity. 2008;29:414–422. doi: 10.1016/j.immuni.2008.06.014. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
112. Lee KH, Holdorf AD, Dustin ML, Chan AC, Allen PM, Shaw AS. T cell receptor signaling precedes immunological synapse formation. Science. 2002;295:1539–1542. doi: 10.1126/science.1067710. [PubMed] [CrossRef] [Google Scholar]
113. Rentero C, Zech T, Quinn CM, Engelhardt K, Williamson D, Grewal T, Jessup W, Harder T, Gaus K. Functional implications of plasma membrane condensation for T cell activation. PLoS ONE. 2008;3:e2262. doi: 10.1371/journal.pone.0002262. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
114. Kovacs B, Parry RV, Ma Z, Fan E, Shivers DK, Freiberg BA, Thomas AK, Rutherford R, Rumbley CA, Riley JL, Finkel TH. Ligation of CD28 by its natural ligand CD86 in the absence of TCR stimulation induces lipid raft polarization in human CD4 T cells. J Immunol. 2005;175:7848–7854. [PubMed] [Google Scholar]
115. Sedwick CE, Morgan MM, Jusino L, Cannon JL, Miller J, Burkhardt JK. TCR, LFA-1, and CD28 play unique and complementary roles in signaling T cell cytoskeletal reorganization. J Immunol. 1999;162:1367–1375. [PubMed] [Google Scholar]
116. Yang H, Reinherz EL. Dynamic recruitment of human CD2 into lipid rafts. Linkage to T cell signal transduction. J Biol Chem. 2001;276:18775–18785. doi: 10.1074/jbc.M009852200. [PubMed] [CrossRef] [Google Scholar]
117. Revy P, Sospedra M, Barbour B, Trautmann A. Functional antigen-independent synapses formed between T cells and dendritic cells. Nat Immunol. 2001;2:925–931. doi: 10.1038/ni713. [PubMed] [CrossRef] [Google Scholar]
118. Salomon B, Bluestone JA. Complexities of CD28/B7: CTLA-4 costimulatory pathways in autoimmunity and transplantation. Annu Rev Immunol. 2001;19:225–252. doi: 10.1146/annurev.immunol.19.1.225. [PubMed] [CrossRef] [Google Scholar]
119. Schneider H, Cai YC, Cefai D, Raab M, Rudd CE. Mechanisms of CD28 signalling. Res Immunol. 1995;146:149–154. doi: 10.1016/0923-2494(96)80248-3. [PubMed] [CrossRef] [Google Scholar]
120. Raab M, Cai YC, Bunnell SC, Heyeck SD, Berg LJ, Rudd CE. p56Lck and p59Fyn regulate CD28 binding to phosphatidylinositol 3-kinase, growth factor receptor-bound protein GRB-2, and T cell-specific protein-tyrosine kinase ITK: implications for T-cell costimulation. Proc Natl Acad Sci USA. 1995;92:8891–8895. doi: 10.1073/pnas.92.19.8891. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
121. Holdorf AD, Green JM, Levin SD, Denny MF, Straus DB, Link V, Changelian PS, Allen PM, Shaw AS. Proline residues in CD28 and the Src homology (SH)3 domain of Lck are required for T cell costimulation. J Exp Med. 1999;190:375–384. doi: 10.1084/jem.190.3.375. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
122. Andres PG, Howland KC, Dresnek D, Edmondson S, Abbas AK, Krummel MF. CD28 signals in the immature immunological synapse. J Immunol. 2004;172:5880–5886. [PubMed] [Google Scholar]
123. Huse M, Klein LO, Girvin AT, Faraj JM, Li QJ, Kuhns MS, Davis MM. Spatial and temporal dynamics of T cell receptor signaling with a photoactivatable agonist. Immunity. 2007;27:76–88. doi: 10.1016/j.immuni.2007.05.017. [PubMed] [CrossRef] [Google Scholar]
124. Bunnell SC, Hong DI, Kardon JR, Yamazaki T, McGlade CJ, Barr VA, Samelson LE. T cell receptor ligation induces the formation of dynamically regulated signaling assemblies. J Cell Biol. 2002;158:1263–1275. doi: 10.1083/jcb.200203043. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
125. Van Komen JS, Mishra S, Byrum J, Chichili GR, Yaciuk JC, Farris AD, Rodgers W. Early and dynamic polarization of T cell membrane rafts and constituents prior to TCR stop signals. J Immunol. 2007;179:6845–6855. [PubMed] [Google Scholar]
126. Bagatolli LA, Gratton E, Fidelio GD. Water dynamics in glycosphingolipid aggregates studied by laurdan fluorescence. Biophys J. 1998;75:331–341. doi: 10.1016/S0006-3495(98)77517-4. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
127. Bagatolli LA, Gratton E. A correlation between lipid domain shape and binary phospholipid mixture composition in free standing bilayers: a two-photon fluorescence microscopy study. Biophys J. 2000;79:434–447. doi: 10.1016/S0006-3495(00)76305-3. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
128. Gaus K, Gratton E, Kable EP, Jones AS, Gelissen I, Kritharides L, Jessup W. Visualizing lipid structure and raft domains in living cells with two-photon microscopy. Proc Natl Acad Sci USA. 2003;100:15554–15559. doi: 10.1073/pnas.2534386100. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
129. Goswami D, Gowrishankar K, Bilgrami S, Ghosh S, Raghupathy R, Chadda R, Vishwakarma R, Rao M, Mayor S. Nanoclusters of GPI-anchored proteins are formed by cortical actin-driven activity. Cell. 2008;135:1085–1097. doi: 10.1016/j.cell.2008.11.032. [PubMed] [CrossRef] [Google Scholar]

Articles from Cellular and Molecular Life Sciences: CMLS are provided here courtesy of Springer

-