首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Structural Organization of Murine Intracisternal A Particles   总被引:9,自引:8,他引:1       下载免费PDF全文
Isolated murine intracisternal A particles have a distinctive set of structural properties as compared to the recognized oncogenic ribonucleic acid tumor viruses. A particles consist of two closely approximated concentric shells. The outer shell is largely, but not entirely, susceptible to treatment with detergents such as deoxycholate and Triton X-100. The inner shell is resistant to these agents and is stabilized against solubilization by sodium dodecyl sulfate as a result of disulfide bonding between its constituent proteins. The major A particle structural protein (molecular weight approximately 70,000) is contained in the inner shell. A protein component with a molecular weight of about 80,000 may be preferentially concentrated in sites of firm attachment between the inner and outer shells.  相似文献   

2.
We show that a comprehensive set of 16 peroxisomal membrane proteins (PMPs) encompassing all types of membrane topologies first target to the endoplasmic reticulum (ER) in Saccharomyces cerevisiae. These PMPs insert into the ER membrane via the protein import complexes Sec61p and Get3p (for tail-anchored proteins). This trafficking pathway is representative for multiplying wild-type cells in which the peroxisome population needs to be maintained, as well as for mutant cells lacking peroxisomes in which new peroxisomes form after complementation with the wild-type version of the mutant gene. PMPs leave the ER in a Pex3p-Pex19p–dependent manner to end up in metabolically active peroxisomes. These results further extend the new concept that peroxisomes derive their basic framework (membrane and membrane proteins) from the ER and imply a new functional role for Pex3p and Pex19p.  相似文献   

3.
Saccharomyces cerevisiae and mammals concerning the mechanisms of the translocation step and discuss the roles of the proteins implicated in this process. Received: 5 June 1996/Revised: 20 September 1996  相似文献   

4.
Mammalian cytosolic Hsp110 family, in concert with the Hsc70:J-protein complex, functions as a disaggregation machinery to rectify protein misfolding problems. Here we uncover a novel role of this machinery in driving membrane translocation during viral entry. The non-enveloped virus SV40 penetrates the endoplasmic reticulum (ER) membrane to reach the cytosol, a critical infection step. Combining biochemical, cell-based, and imaging approaches, we find that the Hsp110 family member Hsp105 associates with the ER membrane J-protein B14. Here Hsp105 cooperates with Hsc70 and extracts the membrane-penetrating SV40 into the cytosol, potentially by disassembling the membrane-embedded virus. Hence the energy provided by the Hsc70-dependent Hsp105 disaggregation machinery can be harnessed to catalyze a membrane translocation event.  相似文献   

5.
An RNA-dependent DNA polymerase associated with intracisternal A particles has been characterized. The enzyme required Mg(2+) or Mn(2+), dithiothreitol and the presence of all four deoxyribonucleoside triphosphates for the expression of maximal activity. Sensitivity of the endogenous RNA-dependent DNA polymerase activity to a low concentration of pancreatic ribonuclease in the presence of a high concentration of NaCl suggested that the enzyme might be utilizing the A particle endogenous RNA as template. Evidence in support of this was provided by analyses of early and late DNA products of the endogenous reaction by Cs(2)SO(4) isopycnic gradient centrifugation and hybridization of purified 60 to 70S and 35S RNAs of A particles with the purified DNA product.  相似文献   

6.
Mitochondrial membrane biogenesis and lipid metabolism require phospholipid transfer from the endoplasmic reticulum (ER) to mitochondria. Transfer is thought to occur at regions of close contact of these organelles and to be nonvesicular, but the mechanism is not known. Here we used a novel genetic screen in S. cerevisiae to identify mutants with defects in lipid exchange between the ER and mitochondria. We show that a strain missing multiple components of the conserved ER membrane protein complex (EMC) has decreased phosphatidylserine (PS) transfer from the ER to mitochondria. Mitochondria from this strain have significantly reduced levels of PS and its derivative phosphatidylethanolamine (PE). Cells lacking EMC proteins and the ER–mitochondria tethering complex called ERMES (the ER–mitochondria encounter structure) are inviable, suggesting that the EMC also functions as a tether. These defects are corrected by expression of an engineered ER–mitochondrial tethering protein that artificially tethers the ER to mitochondria. EMC mutants have a significant reduction in the amount of ER tethered to mitochondria even though ERMES remained intact in these mutants, suggesting that the EMC performs an additional tethering function to ERMES. We find that all Emc proteins interact with the mitochondrial translocase of the outer membrane (TOM) complex protein Tom5 and this interaction is important for PS transfer and cell growth, suggesting that the EMC forms a tether by associating with the TOM complex. Together, our findings support that the EMC tethers ER to mitochondria, which is required for phospholipid synthesis and cell growth.  相似文献   

7.
Proteins to be secreted are transported from the endoplasmic reticulum (ER) to the Golgi apparatus. The transport of these proteins requires the localization and activity of proteins that create ER exit sites, coat proteins to collect cargo and to reshape the membrane into a transport container, and address labels—SNARE proteins—to target the vesicles specifically to the Golgi apparatus. In addition some proteins may need export chaperones or export receptors to enable their exit into transport vesicles. ER export factors, SNAREs, and misfolded Golgi-resident proteins must all be retrieved from the Golgi to the ER again. This retrieval is also part of the organellar homeostasis pathway essential to maintaining the identity of the ER and of the Golgi apparatus. In this review, I will discuss the different processes in retrograde transport from the Golgi to the ER and highlight the mechanistic insights we have obtained in the last couple of years.Proteins that are exposed at the plasma membrane or populate a membrane-bounded organelle are synthesized into the endoplasmic reticulum (ER). In the ER, the folding of these proteins takes place and posttranslational modifications such as N-glycosylation and disulfide bridge formation occur. Upon adopting a suitable, often correct, conformation, proteins destined to locations beyond the ER are concentrated at so-called ER exit sites (ERES) and incorporated into nascent COPII-coated vesicles. These COPII vesicles eventually bud off the ER membrane and are transported to the Golgi (in yeast, Drosophila, and C. elegans) or the ER-Golgi intermediate compartment (in mammalian cells) (Schweizer et al. 1990; Kondylis and Rabouille 2003; Spang 2009; Witte et al. 2011).It is assumed that the vesicle coat is at least partially destabilized through the hydrolysis of GTP by the small GTPase Sar1 (Oka and Nakano 1994; Springer et al. 1999). However, some of the destabilized coat components have to stay on the vesicle until it has reached the Golgi apparatus because coat components participate in the recognition and the tethering process (Barlowe 1997; Cai et al. 2007; Lord et al. 2011; Zong et al. 2012). Subsequently, SNARE proteins on the vesicles (v-SNAREs) zipper up with cognate SNAREs on the Golgi (target SNAREs, t-SNAREs) to drive membrane fusion (Hay et al. 1998; Cao and Barlowe 2000; Parlati et al. 2002). The content of the ER-derived COPII vesicles is thereby released into the lumen of the cis-cisterna of the Golgi apparatus. Most proteins will continue their journey through the Golgi apparatus and encounter further modifications such as extension of the glycosylation tree or lipidation. However, some proteins, especially those involved in the fusion process, i.e., the v-SNAREs or proteins that act as export factors of the ER, such as Vma21, which is essential for export of the correctly folded and assembled V0 sector of the V-ATPase, need to be recycled back to the ER for another round of transport (Ballensiefen et al. 1998; Malkus et al. 2004). Moreover, cis-Golgi proteins are returned to the ER for quality/functional control (Todorow et al. 2000; Sato et al. 2004; Valkova et al. 2011). Finally, some ER-resident proteins, such as the ER Hsp70 chaperone BiP/Kar2, can escape the ER, but are captured at the cis-Golgi by the H/KDEL receptor Erd2 and returned to the ER (Lewis et al. 1990; Semenza et al. 1990; Aoe et al. 1997).Unfortunately, the retrograde transport route is also hijacked by toxins. For example, endocytosed cholera toxin subunit A contains a KDEL sequence and can thereby exploit the system to access the ER (Majoul et al. 1996, 1998). From there, it is retro-translocated into the cytoplasm where it can exert its detrimental function.  相似文献   

8.
Recent findings have suggested that the autophagic isolation membrane (IM) might originate from a domain of the endoplasmic reticulum (ER) called the omegasome. However, the morphological relationships between ER, omegasome, and IM remain unclear. In the present study, we found that hybrid structures composed of a double FYVE domain-containing protein 1 (DFCP1)-positive omegasome and the IM accumulated in Atg3-deficient mouse embryonic fibroblasts (MEFs). Moreover, correlative light and electron microscopy and immunoelectron microscopy revealed that green fluorescent protein (GFP)-tagged DFCP1 was localized on tubular or vesicular elements adjacent to the IM rims. Through detailed morphological analyses, including optimization of a fixation method and electron tomography, we observed a cluster of thin tubular structures between the IM edges and ER, part of which were continuous with IM and/or ER. The formation of these thin tubular clusters was observed in several cell lines and MEFs deficient for Atg5, Atg7, or Atg16L1 but not in FIP200-deficient cells, suggesting that they were relevant to the earlier events in autophagosome formation. Taken together, our findings indicate that these tubular profiles represent a part of the omegasome that links the ER with the IM.  相似文献   

9.
10.
Topogenic determinants that direct protein topology at the endoplasmic reticulum membrane usually function with high fidelity to establish a uniform topological orientation for any given polypeptide. Here we show, however, that through the coupling of sequential translocation events, native topogenic determinants are capable of generating two alternate transmembrane structures at the endoplasmic reticulum membrane. Using defined chimeric and epitope-tagged full-length proteins, we found that topogenic activities of two C-trans (type II) signal anchor sequences, encoded within the seventh and eighth transmembrane (TM) segments of human P-glycoprotein were directly coupled by an inefficient stop transfer (ST) sequence (TM7b) contained within the C-terminus half of TM7. Remarkably, these activities enabled TM7 to achieve both a single- and a double-spanning TM topology with nearly equal efficiency. In addition, ST and C-trans signal anchor activities encoded by TM8 were tightly linked to the weak ST activity, and hence topological fate, of TM7b. This interaction enabled TM8 to span the membrane in either a type I or a type II orientation. Pleiotropic structural features contributing to this unusual topogenic behavior included 1) a short, flexible peptide loop connecting TM7a and TM7b, 2) hydrophobic residues within TM7b, and 3) hydrophilic residues between TM7b and TM8.  相似文献   

11.
The transport of lipids from their synthesis site at the endoplasmic reticulum (ER) to different target membranes could be mediated by both vesicular and nonvesicular transport mechanisms. Nonvesicular lipid transport appears to be the major transport route of certain lipid species, and could be mediated by either spontaneous lipid transport or by lipid-transfer proteins (LTPs). Although nonvesicular lipid transport has been extensively studied for more than four decades, its underlying mechanism, advantage and regulation, have not been fully explored. In particular, the function of LTPs and their involvement in intracellular lipid movement remain largely controversial. In this article, we describe the pathways by which lipids are synthesized at the ER and delivered to different cellular membranes, and discuss the role of LTPs in lipid transport both in vitro and in intact cells.The endoplasmic reticulum (ER) is a large interconnected membrane network that plays a major role in lipid biosynthesis in eukaryotic cells (Borgese et al. 2006). Newly synthesized lipids at the ER are then delivered to different cellular membranes or organelles, each of which shows unique lipid and protein composition and executes distinct cellular function (Holthuis et al. 2003). The transport of lipids from the ER can be mediated by both vesicular and nonvesicular transport mechanisms. Vesicular transport, as opposed to nonvesicular lipid transport, requires metabolic energy, intact cytoskeleton, and connection to the vesicular transport machinery (Kaplan and Simoni 1985a; Voelker 1990; Vance et al. 1991). Although vesicular lipid transport mediates the bulk transport of many lipids, increasing lines of evidence suggest that nonvesicular lipid transport is the major transport route for certain lipid types (Lev 2010). Nonvesicular lipid transport between membranes could be mediated by spontaneous lipid transport, in which a lipid monomer is diffused through the cytosol from a donor to the acceptor membrane. Given that most cellular lipids are highly hydrophobic, their diffusion through an aqueous phase is very slow and insufficient to support substantial transport of most lipids (Jones and Thompson 1989; Mesmin and Maxfield 2009). Nevertheless, spontaneous lipid transport can be greatly facilitated at membrane contact sites (MCSs) (Levine 2004; Holthuis and Levine 2005) and/or by lipid-transfer proteins (LTPs) (Lev 2010). MCSs are defined as small cytosolic gaps of 10–20 nm between the ER membranes and virtually all cellular organelles (Levine 2004; Lebiedzinska et al. 2009), whereas LTPs are intracellular proteins that can carry a lipid monomer in a hydrophobic pocket and transfer it between membranes through an aqueous phase. LTPs were initially discovered as soluble factors that accelerate the exchange or net transfer of different lipid species between membranes in vitro (Wirtz and Zilversmit 1968). Subsequently, many LTPs have been isolated, cloned, and crystallized. LTPs have been identified in all eukaryotes, in plants, and in bacteria, and according to their sequence and structure similarity have been subdivided into different protein families including SEC14, PITP (phosphatidylinositol-transfer protein), START (StAR-related lipid transfer), GLTP (glycolipid transfer protein), SCP-2 (nonspecific LTPs), and OSBP (oxysterol-binding protein)/ORP (OSBP-related proteins) (D’Angelo et al. 2008). In general, LTPs show specificity for one or more lipid types, and may contain only a single lipid-transfer domain (LTD), or additional structural domains with varying functions (Lev 2010). Over the past 40 years, LTPs have been extensively studied and the major principles of their action mode have been established from both biophysical measurements in vitro and structural data (Lev 2010). Nevertheless, the precise function of LTPs in intact cells remains controversial and a subject of an active field of research. In this article, we briefly describe how lipids are synthesized in the ER and delivered to different target membranes, and discuss how LTPs influence lipid transport in vitro and in intact cells.  相似文献   

12.
We observed that the nonfusogenic mouse hepatitis virus (MHV) strain MHV-2 reached a titer of ∼2 log10 higher than that of the fusogenic strain A59 in astrocytoma DBT cells. To determine whether the spike protein is responsible for the difference, a recombinant virus, Penn-98-1, that contains the A59 genome with a spike from MHV-2 was used to infect DBT cells. Results showed that Penn-98-1 behaved like MHV-2, thus establishing a role for the spike protein in viral growth. The inverse correlation between viral fusogenicity and growth was further established in four different cell types and with a fusogenic mutant, the S757R mutant, derived from isogenic Penn-98-1. While both A59 and Penn-98-1 entered cells at similar levels, viral RNA and protein syntheses were significantly delayed for A59. Interestingly, when the genomic RNAs were delivered directly into the cells via transfection, the levels of gene expression for these viruses were similar. Furthermore, cell fractionation experiments revealed that significantly more genomic RNAs for the nonfusogenic MHVs were detected in the endoplasmic reticulum (ER) within the first 2 h after infection than for the fusogenic MHVs. Pretreatment of Penn-98-1 with trypsin reversed its properties in syncytium formation, virus production, and genome transport to the ER. These findings identified a novel role for the spike protein in regulating the uncoating and delivery of the viral genome to the ER after internalization.Murine coronavirus mouse hepatitis virus (MHV) is a member of the family Coronaviridae. It is an enveloped, positive-strand-RNA virus. The viral envelope contains three or four structural proteins, depending on the virus strain (21). The spike (S) protein is a glycoprotein with a molecular mass of approximately 180 kDa. For some MHV strains, such as JHM and A59, the S protein is cleaved by a furin-like proteinase into two subunits, the amino-terminal S1 and the carboxyl-terminal S2. The S1 subunit is thought to form the globular head of the spike and is responsible for the initial attachment of the virus to the receptor on the cell surface. The S2 subunit, which forms the stalk portion of the spike and which anchors the S protein to the viral envelope, facilitates the fusion between the viral envelope and the cell membrane and cell-cell fusion (4, 7, 20, 25, 39). In contrast, the S protein of some other MHV strains, such as MHV-2, does not undergo cleavage and usually does not cause cell-cell fusion (15, 34). It appears that the cleavability of the MHV S protein is associated usually, though not always, with its fusogenicity (10, 36). It has been suggested that the fusogenicity of the S protein may determine the route of virus entry, i.e., via direct fusion with plasma membranes or following endocytosis (11, 34), although the mechanism for virus-induced cell-cell fusion may differ from that for virus-cell fusion during entry (8). The S protein also elicits the induction of neutralizing antibodies and cell-mediated immunity in infected hosts (3). It is therefore an important determinant for viral infectivity, pathogenicity, and virulence (2, 5, 31, 38). The hemagglutinin-esterase (HE) protein is present only in certain MHV strains (22, 42) and may play a role in viral pathogenesis (44, 45). The small envelope (E) protein and the membrane (M) protein play a key role in virus assembly (40). The nucleocapsid (N) protein is a phosphoprotein of approximately 50 kDa and is associated with the RNA genome to form the nucleocapsid inside the envelope (21, 37).Infection of host cells by MHV is mediated through the interaction between the S protein and the cellular receptors that are members of the carcinoembryonic antigen (CEA) family of the immunoglobulin superfamily (9). This interaction then triggers fusion between the viral envelope and the plasma membrane or the endosomal membrane, the latter of which follows receptor-mediated endocytosis, thus allowing the nucleocapsid to deliver into the cytoplasm. Direct entry from the plasma membrane appears to be the predominant route for most MHV strains (19, 28), although entry by some mutant MHVs, such as OBLV60 and MHV-2, is low pH dependent, i.e., via endocytosis (11, 34). However, nothing is known about how the genomic RNA is transported to the rough endoplasmic reticulum (ER) for translation. Once on the ER, the viral genomic RNA is translated into a polymerase polyprotein from the 5′-end two open reading frames (two-thirds of the genome) via ribosomal frameshifting. The polymerase polyproteins in turn synthesize genomic and multiple species of subgenomic mRNAs. These mRNAs are then translated into nonstructural and structural proteins, the latter of which are essential for generation of progeny viruses.MHV can infect rodents, causing hepatitis, enteritis, nephritis, and central nervous system diseases. In the mouse central nervous system, some MHV strains, such as JHM and A59, are neurovirulent, causing acute encephalitis and chronic demyelination (1, 13), while others, such as MHV-2, exhibit extremely low neurovirulence, causing only meningitis without apparent encephalitis and demyelination (6, 16, 41). Extensive mutagenesis studies in combination with targeted RNA recombination have identified that the S protein is the major determinant of MHV pathogenicity in animals, although other viral genes also appear to modulate viral pathogenicity (17, 32). For example, the recombinant MHV Penn-98-1, which contains the S protein of MHV-2 in an A59 genome background, causes acute meningoencephalitis similar to that caused by A59 but does not cause demyelination similar to that observed for MHV-2 (6). It has also been shown that the amounts of antigen staining and necrosis in the liver correlate with the viral titer, which is determined largely by the S protein (29). However, how the S protein affects viral titer in cell culture and in animals is not known.In the present study, we initially observed that the levels of production of infectious viruses in an astrocytoma DBT cell line were markedly different among three MHV strains. Using the recombinant MHV Penn-98-1 and its isogenic S757R mutant, we further established that the S protein is responsible for the observed difference. The difference in virus production between A59 and Penn-98-1 was detected as early as 4 to 6 h postinfection (p.i.) and likely occurred during the early stages of the virus life cycle but after virus internalization. Interestingly, when the genomic RNAs were delivered directly into the cells via transfection, the levels of gene expression for these viruses were similar. Furthermore, cell fractionation experiments revealed that significantly more genomic RNAs for nonfusogenic MHVs were delivered to the ER within the first 2 h after infection than for fusogenic MHVs. These results demonstrate that the spike protein of MHV can regulate the intracellular transport of the viral genome to the ER following internalization. To our knowledge, this is the first study identifying a role for a coronavirus S protein in genome delivery in addition to its well-established role in receptor binding and virus-cell and cell-cell fusions during infection.  相似文献   

13.
The mechanisms by which receptors guide intracellular virus transport are poorly characterized. The murine polyomavirus (Py) binds to the lipid receptor ganglioside GD1a and traffics to the endoplasmic reticulum (ER) where it enters the cytosol and then the nucleus to initiate infection. How Py reaches the ER is unclear. We show that Py is transported initially to the endolysosome where the low pH imparts a conformational change that enhances its subsequent ER-to-cytosol membrane penetration. GD1a stimulates not viral binding or entry, but rather sorting of Py from late endosomes and/or lysosomes to the ER, suggesting that GD1a binding is responsible for ER targeting. Consistent with this, an artificial particle coated with a GD1a antibody is transported to the ER. Our results provide a rationale for transport of Py through the endolysosome, demonstrate a novel endolysosome-to-ER transport pathway that is regulated by a lipid, and implicate ganglioside binding as a general ER targeting mechanism.  相似文献   

14.
Transmembrane proteins are synthesized and folded in the endoplasmic reticulum (ER), an interconnected network of flattened sacs or tubes. Up to now, this organelle has eluded a detailed analysis of the dynamics of its constituents, mainly due to the complex three-dimensional morphology within the cellular cytosol, which precluded high-resolution, single-molecule microscopy approaches. Recent evidences, however, pointed out that there are multiple interaction sites between ER and the plasma membrane, rendering total internal reflection microscopy of plasma membrane proximal ER regions feasible. Here we used single-molecule fluorescence microscopy to study the diffusion of the human serotonin transporter at the ER and the plasma membrane. We exploited the single-molecule trajectories to map out the structure of the ER close to the plasma membrane at subdiffractive resolution. Furthermore, our study provides a comparative picture of the diffusional behavior in both environments. Under unperturbed conditions, the majority of proteins showed similar mobility in the two compartments; at the ER, however, we found an additional 15% fraction of molecules moving with 25-fold faster mobility. Upon degradation of the actin skeleton, the diffusional behavior in the plasma membrane was strongly influenced, whereas it remained unchanged in the ER.Live-cell microscopy and three-dimensional electron tomography has boosted our understanding of endoplasmic reticulum (ER) dynamics and morphology. Proteins have been identified which regulate the formation of cisternae versus tubelike membranes, and the contacts between ER and the various cellular organelles have been studied in detail (1). Little information, however, is available when it comes to protein dynamics and organization within the ER membrane. Its complex three-dimensional topology hampers standard diffraction-limited fluorescence microscopy approaches: in fluorescence recovery after photobleaching, for example, the obtained diffusion coefficients can be several-folds off, if the ER morphology is not correctly taken into account (2). A method is therefore needed which allows for resolving molecular movements on length scales below the typical dimensions of the ER structures.In principle, single-molecule tracking would provide the required spatial resolution due to the high precision in localizing the moving point emitters: localization errors of <40 nm can be easily achieved (3). This technique has given rise to multiple studies, in which the paths of the diffusing objects were used to make conclusions on the properties of the environment; particularly, the plasma membrane has become a favorite target for such investigations, yielding precise determinations of the diffusion coefficients of a variety of membrane proteins or lipids (4).Here, we report what is, to our knowledge, the first application of single-molecule tracking for a comparative study of the diffusion dynamics of a membrane protein at the ER versus the plasma membrane. As the protein of interest, we chose the human serotonin transporter (SERT): it is a polytopic membrane protein containing 12 transmembrane domains, with both C- and N-termini residing in the cytoplasm. Stable SERT oligomers of various degrees were observed to coexist in the plasma membrane (5). Functionally, SERT (6) is a pivotal element in shaping serotonergic neurotransmission: SERT-mediated high-affinity uptake of released serotonin clears the synaptic cleft and supports refilling of vesicular stores (7). Wild-type SERT (SERT-wt) is efficiently targeted to the presynaptic plasma membrane, whereas the truncation of its C-terminus (SERT-ΔC30) retains the mutant protein in the ER (8). The N-terminal mGFP- and eYFP-fusion constructs of the two versions of SERT thus allowed us to specifically address SERT located at the ER (eYFP-SERT-ΔC30) or at the plasma membrane (mGFP-SERT-wt (7)).Our experiments were performed at 37°C on proteins heterologously expressed in CHO cells. Total internal reflection (TIR) illumination afforded a reduction in background fluorescence and allowed for selective imaging of single mGFP-SERT-wt molecules at the cells’ plasma membrane or single eYFP-SERT-ΔC30 molecules at plasma membrane-proximal ER (Fig. 1 and see the Supporting Material). TIR was particularly crucial for single-molecule imaging of the ER-retained mutant, where out-of-focus background would surpass the weak single-molecule signals in epi-illumination.Open in a separate windowFigure 1Schematics of the plasma membrane (PM) and a part of the ER containing mGFP-SERT-wt or the ER-retained eYFP-SERT-ΔC30 mutant, respectively. Both can be excited by total internal reflection fluorescence (TIRF) excitation. Experiments were carried out either on cells expressing mGFP-SERT-wt or eYFP-SERT-ΔC30.For both mutants, the majority of molecules were mobile: in fluorescence-recovery-after-photobleaching experiments we observed a mobile fraction of 82 ± 8% for mGFP-SERT-wt and 91 ± 4% for eYFP-SERT-ΔC30. For single-molecule tracking, the high surface density of signals was reduced by completely photobleaching a rectangular part of the cell in epi-illumination; after a brief recovery period, a few single-molecule signals had entered the bleached area and could be monitored and tracked at high signal/noise using TIR excitation. Samples were illuminated stroboscopically for till = 2 ms, and movies of 500 frames were recorded with a delay of tdel = 6 ms; the short delay times ensured that even rapidly diffusing molecules hardly reached the borders of the ER tubes between two consecutive frames. This illumination protocol was run for 20 times per cell, yielding ∼2500 trajectories per cell.The single-molecule localizations were first used to map those areas that are accessible to the diffusing proteins. eYFP-SERT-ΔC30 showed distinct hotspots, representing plasma membrane-proximal ER, excitable by the evanescent field (Fig. 2 A). These hotspots hardly moved within the timescale of an experiment (tens of minutes, see Fig. S1 in the Supporting Material); indeed, remarkable ER stability was previously observed using superresolution microscopy (9). In contrast, a rather homogeneous distribution was observed for mGFP-SERT-wt in the plasma membrane (Fig. 2 B).Open in a separate windowFigure 2Superresolution and tracking data at the ER and the plasma membrane. Superresolution images are shown for the ER-retained SERT mutant eYFP-SERT-ΔC30 (A) and for mGFP-SERT-wt in the plasma membrane (B). (C and D) Diffusion coefficients of eYFP-SERT-ΔC30 (C) and mGFP-SERT-wt (D) are shown as normalized histograms before (blue) and after (red) Cytochalasin D treatment. Data were fitted by Gaussian mobility distributions (see Table S1 in the Supporting Material for the fit results).Next, we compared the mobility of the observed proteins. Single-molecule localizations were linked to trajectories as described in Gao and Kilfoil (10), and the apparent diffusion coefficient, D, of each molecule was estimated from the first two points of the mean-square displacement membrane. The distribution of log10 D showed a pronounced single peak (Fig. 2 D). It could be well fitted by a linear combination of two Gaussian functions, with the major fraction (85%) characterized by Dwt = 0.30 μm2/s; a broad shoulder to the left indicates the presence of proteins that are immobilized during the observation period. In contrast, the mobility of the ER-retained mutant showed a substantially different distribution, containing two clearly visible peaks (Fig. 2 C). We fitted the data with a three-component Gaussian model: the main fraction (82%) behaved similar to SERT at the plasma membrane, with DΔC30 = 0.32 μm2/s. In addition, a large fraction (15%) with high mobility of DΔC30 = 7.8 μm2/s and a minor fraction (3%) with low mobility was observed. The proteins responded as expected to degradation of the actin membrane skeleton (red bars in Fig. 2, C and D): at the plasma membrane, the mobility of mGFP-SERT-wt increased 4.6-fold (mean values), whereas at the ER membrane there was only a minor change for eYFP-SERT-ΔC30 mobility (1.06-fold increase; note that the ER is not connected to actin filaments (11)).The observation of a high mobility subfraction at the ER membrane is surprising. In general, the presence of obstacles—irrespective of whether randomly distributed or clustered, mobile or immobile—reduces the diffusivity of mobile tracers in a membrane (12). It is generally assumed that the high protein density in cell membranes is responsible for the rather low fluidity when compared to synthetic membranes (compare, e.g., Saxton and Jacobson (13) with Weiss et al. (14)). Interestingly, the observed diffusion constant of 7.8 μm2/s is of similar order as the mobility determined for various proteins in synthetic lipid membranes (14). It is thus tempting to hypothesize the presence of extended protein-depleted regions of higher fluidity within the ER membrane; such membrane domains were indeed observed already at the plasma membrane (15). We were also concerned, however, that protein degradation fragments could have contributed to our data: the three-dimensional mobility of an 85-kDa protein is ∼10 μm2/s (16), similar to the high mobility diffusion constant of eYFP-SERT-ΔC30.We tested the two explanations by analyzing the spatial distribution of fast (DΔC30 > 1 μm2/s) versus slow trajectories (DΔC30 < 1 μm2/s) of eYFP-SERT-ΔC30 (Fig. 3). Both types of trajectories clustered in the same regions, and no segregation into ER subdomains was observable at the resolved length scales. This finding—on the one hand—disfavors freely diffusing protein fragments as the origin of the high mobility fraction. On the other hand, it calls for further experiments to identify the origin of the fast and the slow mobility subfraction. Interestingly, when analyzing all eYFP-SERT-ΔC30 trajectories we found that 80% of the molecules showed diffusion confined to domains of 230-nm radius (see Fig. S2). This size is clearly smaller than the lateral extensions of the visible ER regions observed in Fig. 3. The finding indicates domain formation at the ER membrane; domains are averaged out in Fig. 3 due to the long recording times. Note that free diffusion was observed for mGFP-SERT-wt at the plasma membrane (5).Open in a separate windowFigure 3Ripley’s K function analysis of the different mobility fractions in the ER. For the cell presented in Fig. 2, the first position of every slow (D < 1 μm2/s; red) and fast (D > 1 μm2/s; blue) trajectory was plotted in panel A. Contour lines indicate regions of ER attachment to the plasma membrane. In panel B, the point-correlation function L(r)−r is plotted for the slow (red) and fast (blue) fraction. Furthermore, the correlation between fast versus slow is plotted (green). All three curves show a peak at ∼450 nm, which agrees with the extensions of the ER attachment zones.In conclusion, we have shown that single-molecule tracking is feasible for constituents of the ER membrane. We found a surprising diffusion behavior of SERT resulting in the following:
  • 1.A slow fraction showing mobility reminiscent of protein diffusion in the plasma membrane, likely reflecting SERT diffusing in protein-crowded regions of the ER membrane; and
  • 2.A fast fraction showing 25-fold faster diffusion kinetics.
This likely represents diffusion in altered ER membrane environments, possibly of different lipid or protein composition. Given the fact that synthesis of virtually all membrane proteins and most lipids proceeds at the ER membrane, ER heterogeneity at the nanoscale due to the continuous synthesis activity and selection for correct folding appears highly plausible.  相似文献   

15.
In this article, we show that the endoplasmic reticulum (ER) in Arabidopsis thaliana undergoes morphological changes in structure during ER stress that can be attributed to autophagy. ER stress agents trigger autophagy as demonstrated by increased production of autophagosomes. In response to ER stress, a soluble ER marker localizes to autophagosomes and accumulates in the vacuole upon inhibition of vacuolar proteases. Membrane lamellae decorated with ribosomes were observed inside autophagic bodies, demonstrating that portions of the ER are delivered to the vacuole by autophagy during ER stress. In addition, an ER stress sensor, INOSITOL-REQUIRING ENZYME-1b (IRE1b), was found to be required for ER stress–induced autophagy. However, the IRE1b splicing target, bZIP60, did not seem to be involved, suggesting the existence of an undiscovered signaling pathway to regulate ER stress–induced autophagy in plants. Together, these results suggest that autophagy serves as a pathway for the turnover of ER membrane and its contents in response to ER stress in plants.  相似文献   

16.
Transmembrane proteins are synthesized and folded in the endoplasmic reticulum (ER), an interconnected network of flattened sacs or tubes. Up to now, this organelle has eluded a detailed analysis of the dynamics of its constituents, mainly due to the complex three-dimensional morphology within the cellular cytosol, which precluded high-resolution, single-molecule microscopy approaches. Recent evidences, however, pointed out that there are multiple interaction sites between ER and the plasma membrane, rendering total internal reflection microscopy of plasma membrane proximal ER regions feasible. Here we used single-molecule fluorescence microscopy to study the diffusion of the human serotonin transporter at the ER and the plasma membrane. We exploited the single-molecule trajectories to map out the structure of the ER close to the plasma membrane at subdiffractive resolution. Furthermore, our study provides a comparative picture of the diffusional behavior in both environments. Under unperturbed conditions, the majority of proteins showed similar mobility in the two compartments; at the ER, however, we found an additional 15% fraction of molecules moving with 25-fold faster mobility. Upon degradation of the actin skeleton, the diffusional behavior in the plasma membrane was strongly influenced, whereas it remained unchanged in the ER.  相似文献   

17.
The envelope glycoprotein (Env) of human immunodeficiency virus type 1 (HIV-1) has been shown to redirect the site of virus assembly in polarized epithelial cells. To test whether localization of the glycoprotein exclusively to the endoplasmic reticulum (ER) could redirect virus assembly to that organelle in nonpolarized cells, an ER -retrieval signal was engineered into an epitope-tagged variant of Env. The epitope tag, attached to the C terminus of Env, did not affect the normal maturation and transport of the glycoprotein or the incorporation of Env into virions. The epitope-tagged Env was also capable of mediating syncytium formation and virus entry with a similar efficiency to that of wild-type Env. When the epitope was modified to contain a consensus K(X)KXX ER retrieval signal, however, the glycoprotein was no longer proteolytically processed into its surface and transmembrane subunits and Env could not be detected at the cell surface by biotinylation. Endoglycosidase H analysis revealed that the modified Env was not transported to the Golgi apparatus. Immunofluorescent staining patterns were also consistent with the exclusion of Env from the Golgi. As expected, cells expressing the modified Env failed to form syncytia with CD4+ permissive cells. Despite this tight localization of Env to the ER, when the modified Env was expressed in the context of virus, virions continued to be produced efficiently from the plasma membrane of transfected cells. However, these virions contained no detectable glycoprotein and were noninfectious. Electron microscopy revealed virus budding from the plasma membrane of these cells, but no virus was seen assembling at the ER membrane and no assembled virions were found within the cell. These results suggest that the accumulation of Env in an intracellular compartment is not sufficient to redirect the assembly of HIV Gag in nonpolarized cells.  相似文献   

18.
In biogenesis of membrane proteins on the endoplasmic reticulum, a protein-conducting channel called the translocon functions in both the membrane translocation of lumenal domains and the integration of transmembrane segments. Here we analyzed the environments of polypeptide chains during the processes by water-dependent alkylation of N-ethylmaleimide at site-directed Cys residues. Using the technique, the region embedded in the hydrophobic portion of the membrane within a signal-anchor sequence and its shortening by insertion of a Pro residue could be detected. When translocation of the N-terminal domain of the signal-anchor was arrested by trapping an N-terminally fused affinity tag sequence, the signal-anchor was susceptible to alkylation, indicating that its migration into the hydrophobic environment was also arrested. Furthermore, when the tag sequence was separated from the signal-anchor by insertion of a hydrophilic sequence, the signal-anchor became inaccessible to alkylation even in the N-terminally trapped state. This suggests that membrane integration of the signal-anchor synchronizes with partial translocation of its N-terminal domain. Additionally, in an integration intermediate of a membrane protein, both of the two translocation-arrested hydrophilic chains were in an aqueous environment flanking the translocon, suggesting that the translocon provides the hydrophilic pathway capable of at least two translocating chains.  相似文献   

19.
Hundreds of eukaryotic membrane proteins are anchored to membranes by a single transmembrane domain at their carboxyl terminus. Many of these tail-anchored (TA) proteins are posttranslationally targeted to the endoplasmic reticulum (ER) membrane for insertion by the guided-entry of TA protein insertion (GET) pathway. In recent years, most of the components of this conserved pathway have been biochemically and structurally characterized. Get3 is the pathway-targeting factor that uses nucleotide-linked conformational changes to mediate the delivery of TA proteins between the GET pretargeting machinery in the cytosol and the transmembrane pathway components in the ER. Here we focus on the mechanism of the yeast GET pathway and make a speculative analogy between its membrane insertion step and the ATPase-driven cycle of ABC transporters.The mechanism of membrane protein insertion into the endoplasmic reticulum (ER) has been extensively studied for many years (Shao and Hegde 2011). From this work, the signal recognition particle (SRP)/Sec61 pathway has emerged as a textbook example of a cotranslational membrane insertion mechanism (Grudnik et al. 2009). The SRP binds a hydrophobic segment (either a cleavable amino-terminal signal sequence or a transmembrane domain) immediately after it emerges from the ribosomal exit tunnel. This results in a translational pause that persists until SRP engages its receptor in the ER and delivers the ribosome-nascent chain complex to the Sec61 channel. Last, the Sec61 channel enables protein translocation into the ER lumen along with partitioning of hydrophobic transmembrane domains into the lipid bilayer through the Sec61 lateral gate (Rapoport 2007).Approximately 5% of all eukaryotic membrane proteins have an ER targeting signal in a single carboxy-terminal transmembrane domain that emerges from the ribosome exit tunnel following completion of protein synthesis and is not recognized by the SRP (Stefanovic and Hegde 2007). Nonetheless, because hydrophobic peptides in the cytoplasm are prone to aggregation and subject to degradation by quality control systems (Hessa et al. 2011), these tail-anchored (TA) proteins still have to be specifically recognized, shielded from the aqueous environment, and guided to the ER membrane for insertion. In the past five years, the guided-entry of TA proteins (GET) pathway has come to prominence as the major machinery for performing these tasks and the enabler of many key cellular processes mediated by TA proteins including vesicle fusion, membrane protein insertion, and apoptosis. This research has rapidly yielded biochemical and structural insights (and2)2) into many of the GET pathway components (Hegde and Keenan 2011; Chartron et al. 2012a; Denic 2012). In particular, Get3 is an ATPase that uses metabolic energy to bridge recognition of TA proteins by upstream pathway components with TA protein recruitment to the ER for membrane insertion. However, the precise mechanisms of nucleotide-dependent TA protein binding to Get3 and how the GET pathway inserts tail anchors into the membrane are still poorly understood. Here, we provide an overview of the budding yeast GET pathway with emphasis on mechanistic insights that have come from structural studies of its membrane-associated steps and make a speculative juxtaposition with the ABC transporter mechanism.

Table 1.

A catalog of GET pathway component structures
ComponentRole in the pathwayPDB ID
Sgt2Component of the pretargeting complex that delivers TA proteins to Get3; dimer interacts with Get4/Get5, contains TPR repeats that interact with Hsps3SZ7
Get5Component of the pretargeting complex that delivers TA proteins to Get3; dimer interacts with Get4 via amino-terminal domain and with Sgt2 via its ubiquitin-like domain2LNZ
3VEJ
2LO0
Get4Component of the pretargeting complex that delivers TA proteins to Get3; interacts with Get3 via amino-terminal domain and with Get4 via carboxy-terminal domain3LPZ
3LKU
3WPV
Get3ATPase that binds the TA protein; dimer interacts with the pretargeting complex in the cytosol, and with Get1/2 at the ER membraneTable 2
Get1ER receptor for Get3; integral ER membrane
protein, three TMDs; forms a complex with Get2
3SJA, 3SJB
3SJC, 3ZS8
3VLC, 3B2E
Get2ER receptor for Get3; integral ER membrane
protein, three TMDs; forms a complex with Get1
3SJD
3ZS9
Open in a separate windowTA, tail anchored; TPR, tetratricopeptide repeat; TMDs, transmembrane domains.

Table 2.

An itemized list of published Get3 structures with associated nucleotides and conformation nomenclature
OrganismNucleotideConformationPDB IDReferences
Get3
Schizosaccharomyces pombeNoneOpen2WOOMateja et al. 2009
Saccharomyces cerevisiaeNoneOpen3H84Hu et al. 2009
3A36Yamagata et al. 2010
Aspergillus fumigatusADPOpen3IBGSuloway et al. 2009
S. cerevisiaeADPOpen3A37Yamagata et al. 2010
Debaryomyces hanseniiADPClosed3IO3Hu et al. 2009
Chaetomium thermophilumAMPPNP-Mg2+Closed3IQWBozkurt et al. 2009
C. thermophilumADP-Mg2+Closed3IQXBozkurt et al. 2009
S. cerevisiaeADP•AlF4-Mg2+Fully closed2WOJMateja et al. 2009
Methanothermobacter thermautotrophicusADP•AlF4-Mg2+Fully closed3ZQ6Sherill et al. 2011
Methanococcus jannaschiiADP•AlF4-Mg2+Tetrameric3UG6Suloway et al. 2012
3UG7
Get3/Get2cyto
S. cerevisiaeADP-Mg2+Closed3SJDStefer et al. 2011
S. cerevisiaeADP•AlF4-Mg2+Closed3ZS9Mariappan et al. 2011
Get3/Get1cyto
S. cerevisiaeNoneSemiopen3SJCStefer et al. 2011
S. cerevisiaeADPSemiopen3VLCKubota et al. 2012
S. cerevisiaeNoneOpen3SJAStefer et al. 2011
3SJBStefer et al. 2011
3ZS8Mariappan et al. 2011
ADPOpen3B2EKubota et al. 2012
Open in a separate windowADP, adenosine diphosphate.  相似文献   

20.
The cytoplasmic replication of positive-sense RNA viruses is associated with a dramatic rearrangement of host cellular membranes. These virus-induced changes result in the induction of vesicular structures that envelop the virus replication complex (RC). In this study, we have extended our previous observations on the intracellular location of West Nile virus strain Kunjin virus (WNVKUN) to show that the virus-induced recruitment of host proteins and membrane appears to occur at a pre-Golgi step. To visualize the WNVKUN replication complex, we performed three-dimensional (3D) modeling on tomograms from WNVKUN replicon-transfected cells. These analyses have provided a 3D representation of the replication complex, revealing the open access of the replication complex with the cytoplasm and the fluidity of the complex to the rough endoplasmic reticulum. In addition, we provide data that indicate that a majority of the viral RNA species housed within the RC is in a double-stranded RNA (dsRNA) form.West Nile virus (WNV) belongs to the Flaviviridae, which is a large family of enveloped, positive-strand RNA viral pathogens that are responsible for causing severe disease and mortality in humans and animals each year. The Australian WNV strain Kunjin virus (WNVKUN) is a relatively low-pathogenic virus that is closely related to the pathogenic WNV strain New York 99 (WNVNY99), the causative agent of the 1999 epidemic of encephalitis in New York City (11).It has become increasingly known that the replication of most, if not all, positive-sense RNA viruses, whether they infect plants, insects, or humans, is associated with dramatic membrane alterations resulting in the formation of membranous microenvironments that facilitate efficient virus replication. In most cases the induced membrane structures house the actively replicating viral RNA and comprise 70- to 100-nm membrane “vesicles” (sometimes referred to as spherules). Although this distinct morphology is shared across virus families, the cellular origins of these membranes is diverse: the endoplasmic reticulum (ER), mitochondria, peroxisomes, and trans-Golgi membranes have been implicated in different viral systems (1, 8, 13, 23, 31, 38, 41, 45). This diversity implies that the processes involved in inducing the membrane vesicles/spherules are shared, rather than the composition of the membrane itself, although the exact purpose for utilizing membranes derived from different cellular compartments is still not completely resolved or understood.The replication of the flavivirus WNVKUN is associated with the induction of morphologically distinct membrane structures that have defined roles during the WNVKUN replication cycle. Three well-defined structures can be seen as large convoluted membranes (CM), paracrystalline arrays (PC), or membrane sacs containing small vesicles, termed vesicle packets (VP) (18, 20, 48). Based on localization studies with viral proteins of specific functions, we observed that components of the virus protease complex (namely, nonstructural protein 3 [NS3] with cofactor NS2B) localize specifically to the CM/PC, whereas viral double-stranded RNA (dsRNA) and the viral RNA-dependent RNA polymerase (RdRp) NS5 localized primarily to VP (20-22, 47, 48). Additionally, we observed that the CM and PC originate from and are modified membranes of the intermediate compartment (IC) and rough endoplasmic reticulum (RER), whereas the VP appear to be derived from trans-Golgi network (TGN) membranes (19). Recently, we have found that the WNVKUN NS4A protein by itself has the intrinsic capacity to induce the CM and PC structures (35), a property also subsequently shown for Dengue virus (DENV) NS4A (29). Additionally, we have shown that upon WNV infection cellular cholesterol and cholesterol-synthesizing proteins are redistributed to the virus-induced membranes and that this redistribution severely disrupted the formation of cholesterol-rich microdomains (23). Furthermore, we have shown that the membranous structures induced during WNV replication provide partial protection of the WNV replication components from the interferon (IFN)-induced antiviral MxA protein, suggesting that the distinct compartmentalization of viral replication and components of the cellular antiviral response may be an evolutionary mechanism by which flaviviruses can protect themselves from host surveillance (6).In this study we focused on three-dimensional (3D) modeling to give insight into the 3D structure of the VP and provide evidence of how these complexes are organized and formed within the RER membrane. These results add valuable information to our understanding of how the WNV replication complex (RC) functions.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号