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.
《Biophysical journal》2020,118(2):403-414
Nanodiscs (NDs) are an excellent alternative to small unilamellar vesicles (SUVs) for studies of membrane protein structure, but it has not yet been shown that membrane proteins are able to spontaneously fold and insert into a solution of freely diffusing NDs. In this article, we present SDS-PAGE differential mobility studies combined with fluorescence, circular dichroism, and ultraviolet resonance Raman spectroscopy to confirm the spontaneous folding of outer membrane protein A (OmpA) into preformed NDs. Folded OmpA in NDs was incubated with Arg-C protease, resulting in the digestion of OmpA to membrane-protected fragments with an apparent molecular mass of ∼26 kDa (major component) and ∼24 kDa (minor component). The OmpA folding yields were greater than 88% in both NDs and SUVs. An OmpA adsorbed intermediate on NDs could be isolated at low temperature and induced to fold via an increase in temperature, analogous to the temperature-jump experiments on SUVs. The circular dichroism spectra of OmpA in NDs and SUVs were similar and indicated β-barrel secondary structure. Further evidence of OmpA folding into NDs was provided by ultraviolet resonance Raman spectroscopy, which revealed the intense 785 cm−1 structural marker for folded OmpA in NDs. The primary difference between folding in NDs and SUVs was the kinetics; the rate of folding was two- to threefold slower in NDs compared to in SUVs, and this decreased rate can tentatively be attributed to the properties of NDs. These data indicate that NDs may be an excellent alternative to SUVs for folding experiments and offer benefits of optical clarity, sample homogeneity, control of ND:protein ratios, and greater stability.  相似文献   

2.
Amyloidosis resulting from the deposition of aggregated protein has been linked to many debilitating degenerative diseases which include most notably Alzheimer's and Parkinson's. The tendency for a protein to alternatively form highly ordered amyloid fibrils is dependent on many biological factors. Mutations, temperature, concentration, translational motion and pH play a pivotal role in inducing fibril aggregate assembly in vitro. The key feature appears to be the need to destabilize the native state structure as a required first step. In this paper we report on the detailed conversion of the death domain of the human Fas-associated death domain, an all α-helical protein with a Greek-key topology, into an all β-sheet amyloid fibril, using a comprehensive range of spectroscopic techniques that provide insight into this process. This transition from α-helical to β-sheet seems to require destabilization but not complete loss of the secondary structure to explore alternative conformations. This is a fascinating transition that supports the hypothesis that all proteins have the innate ability to form a fibril-like structure. Thus, the primary structure can encode two alternative three-dimensional structures: the native, functional state and the β-amyloid state. The Fas-associated death domain does not appear to naturally form amyloid fibrils in vivo. Our results clearly indicate that proteins evolved to avoid amyloid fibril formation because we find that the conditions required for formation in our model system are very specific and far from physiological.  相似文献   

3.
Computational design has been used with mixed success for the design of protein surfaces, with directed evolution heretofore providing better practical solutions than explicit design. Directed evolution, however, requires a tractable high-throughput screen because the random nature of mutation does not enrich for desired traits. Here we demonstrate the successful design of the β-sheet surface of a red fluorescent protein (RFP), enabling control over its oligomerization. To isolate the problem of surface design, we created a hybrid RFP from DsRed and mCherry with a stabilized protein core that allows for monomerization without loss of fluorescence. We designed an explicit library for which 93 of 96 (97%) of the protein variants are soluble, stably fluorescent, and monomeric. RFPs are heavily used in biology, but are natively tetrameric, and creating RFP monomers has proven extremely difficult. We show that surface design and core engineering are separate problems in RFP development and that the next generation of RFP markers will depend on improved methods for core design.  相似文献   

4.
Although the amyloid dye thioflavin-T (ThT) is among the most widely used tools in the study of amyloid fibrils, the mechanism by which ThT binds to fibrils and other β-rich peptide self-assemblies remains elusive. The development of the water-soluble peptide self-assembly mimic (PSAM) system has provided a set of ideal model proteins for experimentally exploring the properties and minimal dye-binding requirements of amyloid fibrils. PSAMs consist of a single-layer β-sheet (SLB) capped by two globular domains, which capture the flat, extended β-sheet features common among fibril-like surfaces. Recently, a PSAM that binds to ThT with amyloid-like affinity (low micromolar Kd) has been designed, and its crystal structure in the absence of bound ThT was determined. This PSAM thus provides a unique opportunity to examine the interactions of ThT with a β-rich structure. Here, we present molecular dynamics simulations of the binding of ThT to this PSAM β-sheet. We show that the primary binding site for ThT is along a shallow groove formed by adjacent Tyr and Leu residues on the β-sheet surface. These simulations provide an atomic-scale rationale for this PSAM's experimentally determined dye-binding properties. Together, our results suggest that an aromatic-hydrophobic groove spanning across four consecutive β-strands represents a minimal ThT binding site on amyloid fibrils. Grooves formed by aromatic-hydrophobic residues on amyloid fibril surfaces may therefore offer a generic mode of recognition for amyloid dyes.  相似文献   

5.
The evolutionarily conserved protein Omp85 is required for outer membrane protein (OMP) assembly in gram-negative bacteria and in mitochondria. Its Escherichia coli homolog, designated BamA, functions with four accessory lipoproteins, BamB, BamC, BamD, and BamE, together forming the β-barrel assembly machinery (Bam). Here, we addressed the composition of this machinery and the function of its components in Neisseria meningitidis, a model organism for outer membrane biogenesis studies. Analysis of genome sequences revealed homologs of BamC, BamD (previously described as ComL), and BamE and a second BamE homolog, Mlp. No homolog of BamB was found. As in E. coli, ComL/BamD appeared essential for viability and for OMP assembly, and it could not be replaced by its E. coli homolog. BamE was not essential but was found to contribute to the efficiency of OMP assembly and to the maintenance of OM integrity. A bamC mutant showed only marginal OMP assembly defects, but the impossibility of creating a bamC bamE double mutant further indicated the function of BamC in OMP assembly. An mlp mutant was unaffected in OMP assembly. The results of copurification assays demonstrated the association of BamC, ComL, and BamE with Omp85. Semi-native gel electrophoresis identified the RmpM protein as an additional component of the Omp85 complex, which was confirmed in copurification assays. RmpM was not required for OMP folding but stabilized OMP complexes. Thus, the Bam complex in N. meningitidis consists of Omp85/BamA plus RmpM, BamC, ComL/BamD, and BamE, of which ComL/BamD and BamE appear to be the most important accessory components for OMP assembly.Membrane-embedded β-barrel proteins are found in the outer membranes (OMs) of gram-negative bacteria, mitochondria, and chloroplasts. Only in recent years have cellular components required for the assembly and insertion of these OM proteins (OMPs) into the OM been identified. Omp85, which was first characterized in Neisseria meningitidis, is the key protein of the OMP assembly machinery (41). The function of Omp85 has been preserved during evolution, not only in gram-negative bacteria (8, 37, 44, 46) but also in mitochondria, where an Omp85 homolog, also known as Tob55 or Sam50, was shown to mediate the assembly of β-barrel proteins into the OM (15, 23, 27). Accordingly, bacterial OMPs are still recognized by the eukaryotic assembly machinery: when expressed in yeast, bacterial OMPs were found to be assembled into the mitochondrial OM in a Tob55-dependent manner (43). Omp85 in Escherichia coli, which was recently renamed BamA, for β-barrel assembly machinery (Bam) component A, is associated with at least four lipoproteins: BamB (formerly known as YfgL), BamC (NlpB), BamD (YfiO), and BamE (SmpA) (32, 46). In E. coli, BamB, BamC, and BamE are not essential, but the phenotypes of deletion mutants suggest that these proteins contribute to the efficiency of OMP assembly. Like BamA, BamD is an essential protein in E. coli (24, 26), involved in OMP assembly (24). These lipoproteins are evolutionarily less well conserved; the mitochondrial Tob55 protein is associated with two accessory proteins, but they do not show any sequence similarity with the lipoproteins of the E. coli Bam complex (14).Besides E. coli, N. meningitidis is one of the major bacterial model organisms for studies of OM assembly. As mentioned above, it was the first organism in which the function of Omp85 was identified (41), and also, the role of an integral OMP, designated LptD (formerly Imp or OstA), in the transport of lipopolysaccharide (LPS) to the cell surface was first established in N. meningitidis (3). With regard to OM biogenesis, N. meningitidis has several features that distinguish it from E. coli. For example, in contrast to E. coli (13), N. meningitidis mutants defective in LPS synthesis or transport are viable (3, 34), and OMPs are assembled perfectly well in such mutants (33). Furthermore, in OMP assembly mutants of E. coli, the periplasmic accumulation of unassembled OMPs is limited due to the induction of the σE extracytoplasmic stress response, which results in the degradation of unfolded OMPs (30) and the inhibition of their synthesis by small regulatory RNAs (20). In contrast, in N. meningitidis, most of the components involved in this response are absent (4), and unassembled OMPs continue to accumulate as periplasmic aggregates when OMP assembly is halted (41). However, the composition of the Bam complex and the role of accessory components in OMP assembly have not so far been studied in this organism. Therefore, to further understand the OMP assembly process in N. meningitidis, we have now analyzed the composition of the Bam complex and addressed the roles of the different components.  相似文献   

6.
The physical interactions that switch the functional state of membrane proteins are poorly understood. Previously, the pH-gating conformations of the β-barrel forming outer membrane protein G (OmpG) from Escherichia coli have been solved. When the pH changes from neutral to acidic the flexible extracellular loop L6 folds into and closes the OmpG pore. Here, we used single-molecule force spectroscopy to structurally localize and quantify the interactions that are associated with the pH-dependent closure. At acidic pH, we detected a pH-dependent interaction at loop L6. This interaction changed the (un)folding of loop L6 and of β-strands 11 and 12, which connect loop L6. All other interactions detected within OmpG were unaffected by changes in pH. These results provide a quantitative and mechanistic explanation of how pH-dependent interactions change the folding of a peptide loop to gate the transmembrane pore. They further demonstrate how the stability of OmpG is optimized so that pH changes modify only those interactions necessary to gate the transmembrane pore.  相似文献   

7.
The assembly of the β-barrel proteins present in the outer membrane (OM) of Gram-negative bacteria is poorly characterized. After translocation across the inner membrane, unfolded β-barrel proteins are escorted across the periplasm by chaperones that reside within this compartment. Two partially redundant chaperones, SurA and Skp, are considered to transport the bulk mass of β-barrel proteins. We found that the periplasmic disulfide isomerase DsbC cooperates with SurA and the thiol oxidase DsbA in the folding of the essential β-barrel protein LptD. LptD inserts lipopolysaccharides in the OM. It is also the only β-barrel protein with more than two cysteine residues. We found that surAdsbC mutants, but not skpdsbC mutants, exhibit a synthetic phenotype. They have a decreased OM integrity, which is due to the lack of the isomerase activity of DsbC. We also isolated DsbC in a mixed disulfide complex with LptD. As such, LptD is identified as the first substrate of DsbC that is localized in the OM. Thus, electrons flowing from the cytoplasmic thioredoxin system maintain the integrity of the OM by assisting the folding of one of the most important β-barrel proteins.  相似文献   

8.
Despite its central role in the protein folding process, the specific mechanism(s) behind β-sheet formation has yet to be determined. For example, whether the nucleation of β-sheets, often containing strands separated in sequence by many residues, is local or not remains hotly debated. Here, we investigate the initial nucleation step of β-sheet formation by performing an analysis of the smallest β-sheets in a non-redundant dataset on the grounds that the smallest sheets, having undergone little growth after nucleation, will be enriched for nucleating characteristics. We find that the residue propensities are similar for small and large β-sheets as are their interstrand pairing preferences, suggesting that nucleation is not primarily driven by specific residues or interacting pairs. Instead, an examination of the structural environments of the two-stranded sheets shows that virtually all of them are contained in single, compact structural modules, or when multiple modules are present, one or both of the chain termini are involved. We, therefore, find that β-nucleation is a local phenomenon resulting either from sequential or topological proximity. We propose that β-nucleation is a result of two opposite factors; that is, the relative rigidity of an associated folding module that holds two stretches of coil close together in topology coupled with sufficient chain flexibility that enables the stretches of coil to bring their backbones in close proximity. Our findings lend support to the hydrophobic zipper model of protein folding (Dill, K. A., Fiebig, K. M., and Chan, H. S. (1993) Proc. Natl. Acad. Sci. U.S.A. 90, 1942–1946). Implications for protein folding are discussed.  相似文献   

9.
The β-barrel assembly machinery (BAM) complex mediates the assembly of β-barrel membrane proteins in the outer membrane. BepA, formerly known as YfgC, interacts with the BAM complex and functions as a protease/chaperone for the enhancement of the assembly and/or degradation of β-barrel membrane proteins. To elucidate the molecular mechanism underlying the dual functions of BepA, its full-length three-dimensional structure is needed. Here, we report the crystal structure of full-length BepA at 2.6-Å resolution. BepA possesses an N-terminal protease domain and a C-terminal tetratricopeptide repeat domain, which interact with each other. Domain cross-linking by structure-guided introduction of disulfide bonds did not affect the activities of BepA in vivo, suggesting that the function of this protein does not involve domain rearrangement. The full-length BepA structure is compatible with the previously proposed docking model of BAM complex and tetratricopeptide repeat domain of BepA.  相似文献   

10.

Background

A key step of anthrax toxin action involves the formation of a protein-translocating pore within the endosomal membrane by the Protective Antigen (PA) moiety. Formation of this transmembrane pore by PA involves interaction of the seven 2β2–2β3 loops of the heptameric precursor to generate a 14-strand transmembrane β barrel.

Methodology/Principal Findings

We examined the effects on pore formation, protein translocation, and cytotoxicity, of mutating two phenylalanines, F313 and F314, that lie at the tip the β barrel, and a third one, F324, that lies part way up the barrel.

Conclusions/Significance

Our results show that the function of these phenylalanine residues is to mediate membrane insertion and formation of stable transmembrane channels. Unlike F427, a key luminal residue in the cap of the pore, F313, F314, and F324 do not directly affect protein translocation through the pore. Our findings add to our knowledge of structure-function relationships of a key virulence factor of the anthrax bacillus.  相似文献   

11.
The severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) envelope (E) protein forms a pentameric ion channel in the lipid membrane of the endoplasmic reticulum Golgi intermediate compartment (ERGIC) of the infected cell. The cytoplasmic domain of E interacts with host proteins to cause virus pathogenicity and may also mediate virus assembly and budding. To understand the structural basis of these functions, here we investigate the conformation and dynamics of an E protein construct (residues 8–65) that encompasses the transmembrane domain and the majority of the cytoplasmic domain using solid-state NMR. 13C and 15N chemical shifts indicate that the cytoplasmic domain adopts a β-sheet-rich conformation that contains three β-strands separated by turns. The five subunits associate into an umbrella-shaped bundle that is attached to the transmembrane helices by a disordered loop. Water-edited NMR spectra indicate that the third β-strand at the C terminus of the protein is well hydrated, indicating that it is at the surface of the β-bundle. The structure of the cytoplasmic domain cannot be uniquely determined from the inter-residue correlations obtained here due to ambiguities in distinguishing intermolecular and intramolecular contacts for a compact pentameric assembly of this small domain. Instead, we present four structural topologies that are consistent with the measured inter-residue contacts. These data indicate that the cytoplasmic domain of the SARS-CoV-2 E protein has a strong propensity to adopt β-sheet conformations when the protein is present at high concentrations in lipid bilayers. The equilibrium between the β-strand conformation and the previously reported α-helical conformation may underlie the multiple functions of E in the host cell and in the virion.  相似文献   

12.
We present molecular dynamics simulation studies of the structural stability of an enclosed loop in the β domain of the Escherichia coli O157:H7 autotransporter EspP. Our investigation revealed that, in addition to its excellent resistance to thermal perturbations, EspP loop 5 (L5) also has remarkable mechanical stability against pulling forces along the membrane norm. These findings are consistent with the experimental report that EspP L5 helps to maintain the permeability barrier in the outer membrane. In contrast to the major secondary structure elements of globular proteins such as ubiquitin, whose resistance to thermal and mechanical perturbations depends mainly on backbone hydrogen bonds and hydrophobic interactions, the structural stability of EspP L5 can be attributed mainly to geometric constraints and side-chain interactions dominated by hydrogen bonds. Examination of B-factors from available high-resolution structures of membrane-embedded β barrels indicates that most of the enclosed loops have stable structures. This finding suggests that loops stabilized by geometric constraints and side-chain interactions might be used more generally to restrict β-barrel channels for various functional purposes.  相似文献   

13.
The endoplasmic reticulum mitochondria encounter structure (ERMES) tethers the ER to mitochondria and contains four structural components: Mmm1, Mdm12, Mdm10, and Mmm2 (Mdm34). The Gem1 protein may play a role in regulating ERMES function. Saccharomyces cerevisiae and Neurospora crassa strains lacking any of Mmm1, Mdm12, or Mdm10 are known to show a variety of phenotypic defects including altered mitochondrial morphology and defects in the assembly of β-barrel proteins into the mitochondrial outer membrane. Here we examine ERMES complex components in N. crassa and show that Mmm1 is an ER membrane protein containing a Cys residue near its N-terminus that is conserved in the class Sordariomycetes. The residue occurs in the ER-lumen domain of the protein and is involved in the formation of disulphide bonds that give rise to Mmm1 dimers. Dimer formation is required for efficient assembly of Tom40 into the TOM complex. However, no effects are seen on porin assembly or mitochondrial morphology. This demonstrates a specificity of function and suggests a direct role for Mmm1 in Tom40 assembly. Mutation of a highly conserved region in the cytosolic domain of Mmm1 results in moderate defects in Tom40 and porin assembly, as well as a slight morphological phenotype. Previous reports have not examined the role of Mmm2 with respect to mitochondrial protein import and assembly. Here we show that absence of Mmm2 affects assembly of β-barrel proteins and that lack of any ERMES structural component results in defects in Tom22 assembly. Loss of N. crassa Gem1 has no effect on the assembly of these proteins but does affect mitochondrial morphology.  相似文献   

14.
The β-barrel assembly machinery (BAM) mediates folding and insertion of β-barrel outer membrane proteins (OMPs) into the outer membrane of Gram-negative bacteria. BAM is a five-protein complex consisting of the β-barrel OMP BamA and lipoproteins BamB, -C, -D, and -E. High resolution structures of all the individual BAM subunits and a BamD-BamC complex have been determined. However, the overall complex architecture remains elusive. BamA is the central component of BAM and consists of a membrane-embedded β-barrel and a periplasmic domain with five polypeptide translocation-associated (POTRA) motifs thought to interact with the accessory lipoproteins. Here we report the crystal structure of a fusion between BamB and a POTRA3–5 fragment of BamA. Extended loops 13 and 17 protruding from one end of the BamB β-propeller contact the face of the POTRA3 β-sheet in BamA. The interface is stabilized by several hydrophobic contacts, a network of hydrogen bonds, and a cation-π interaction between BamA Tyr-255 and BamB Arg-195. Disruption of BamA-BamB binding by BamA Y255A and probing of the interface by disulfide bond cross-linking validate the physiological relevance of the observed interface. Furthermore, the structure is consistent with previously published mutagenesis studies. The periplasmic five-POTRA domain of BamA is flexible in solution due to hinge motions in the POTRA2–3 linker. Modeling BamB in complex with full-length BamA shows BamB binding at the POTRA2–3 hinge, suggesting a role in modulation of BamA flexibility and the conformational changes associated with OMP folding and insertion.  相似文献   

15.
CEL-III is a hemolytic lectin isolated from the sea cucumber Cucumaria echinata. This lectin is composed of two carbohydrate-binding domains (domains 1 and 2) and one oligomerization domain (domain 3). After binding to the cell surface carbohydrate chains through domains 1 and 2, domain 3 self-associates to form transmembrane pores, leading to cell lysis or death, which resembles other pore-forming toxins of diverse organisms. To elucidate the pore formation mechanism of CEL-III, the crystal structure of the CEL-III oligomer was determined. The CEL-III oligomer has a heptameric structure with a long β-barrel as a transmembrane pore. This β-barrel is composed of 14 β-strands resulting from a large structural transition of α-helices accommodated in the interface between domains 1 and 2 and domain 3 in the monomeric structure, suggesting that the dissociation of these α-helices triggered their structural transition into a β-barrel. After heptamerization, domains 1 and 2 form a flat ring, in which all carbohydrate-binding sites remain bound to cell surface carbohydrate chains, stabilizing the transmembrane β-barrel in a position perpendicular to the plane of the lipid bilayer.  相似文献   

16.
X-ray crystallography has been a useful tool in the development of site-directed spin labeling by resolving rotamers of the nitroxide spin-label side chain in a variety of α-helical environments. In this work, the crystal structure of a doubly spin-labeled N8C/K28C mutant of the B1 immunoglobulin-binding domain of protein G (GB1) was solved. The double mutant formed a domain-swapped dimer under crystallization conditions. Two rotameric states of the spin-label were resolved at the solvent-exposed α-helical site, at residue 28; these are in good agreement with rotamers previously reported for helical structures. The second site, at residue 8 on an interior β-strand, shows the presence of three distinct solvent-exposed side-chain rotamers. One of these rotamers is rarely observed within crystal structures of R1 sites and suggests that the H(α) and S(δ) hydrogen bond that is common to α-helical sites is absent at this interior β-strand residue. Variable temperature continuous wave (CW) experiments of the β-strand site showed two distinct components that were correlated to the rotameric states observed in crystallography. Interestingly, the CW data at room temperature could be fit without the use of an order parameter, which is consistent with the lack of the H(α) and S(δ) interaction. Additionally, double electron electron resonance (DEER) spectroscopy was performed on the GB1 double mutant in its monomeric form and yielded a most probable interspin distance of 25 ± 1 ?. In order to evaluate the accuracy of the measured DEER distance, the rotamers observed in the crystal structure of the domain-swapped GB1 dimer were modeled into a high-resolution structure of the wild type monomeric GB1. The distances generated in the resulting GB1 structural models match the most probable DEER distance within ~2 ?. The results are interesting as they indicate by direct experimental measurement that the rotameric states of R1 found in this crystal provide a very close match to the most probable distance measured by DEER.  相似文献   

17.
We expressed the γ-subspecies of protein kinase C (γ-PKC) fused with green fluorescent protein (GFP) in various cell lines and observed the movement of this fusion protein in living cells under a confocal laser scanning fluorescent microscope. γ-PKC–GFP fusion protein had enzymological properties very similar to that of native γ-PKC. The fluorescence of γ-PKC– GFP was observed throughout the cytoplasm in transiently transfected COS-7 cells. Stimulation by an active phorbol ester (12-O-tetradecanoylphorbol 13-acetate [TPA]) but not by an inactive phorbol ester (4α-phorbol 12, 13-didecanoate) induced a significant translocation of γ-PKC–GFP from cytoplasm to the plasma membrane. A23187, a Ca2+ ionophore, induced a more rapid translocation of γ-PKC–GFP than TPA. The A23187-induced translocation was abolished by elimination of extracellular and intracellular Ca2+. TPA- induced translocation of γ-PKC–GFP was unidirected, while Ca2+ ionophore–induced translocation was reversible; that is, γ-PKC–GFP translocated to the membrane returned to the cytosol and finally accumulated as patchy dots on the plasma membrane. To investigate the significance of C1 and C2 domains of γ-PKC in translocation, we expressed mutant γ-PKC–GFP fusion protein in which the two cysteine rich regions in the C1 region were disrupted (designated as BS 238) or the C2 region was deleted (BS 239). BS 238 mutant was translocated by Ca2+ ionophore but not by TPA. In contrast, BS 239 mutant was translocated by TPA but not by Ca2+ ionophore. To examine the translocation of γ-PKC–GFP under physiological conditions, we expressed it in NG-108 cells, N-methyl-d-aspartate (NMDA) receptor–transfected COS-7 cells, or CHO cells expressing metabotropic glutamate receptor 1 (CHO/mGluR1 cells). In NG-108 cells , K+ depolarization induced rapid translocation of γ-PKC–GFP. In NMDA receptor–transfected COS-7 cells, application of NMDA plus glycine also translocated γ-PKC–GFP. Furthermore, rapid translocation and sequential retranslocation of γ-PKC–GFP were observed in CHO/ mGluR1 cells on stimulation with the receptor. Neither cytochalasin D nor colchicine affected the translocation of γ-PKC–GFP, indicating that translocation of γ-PKC was independent of actin and microtubule. γ-PKC–GFP fusion protein is a useful tool for investigating the molecular mechanism of γ-PKC translocation and the role of γ-PKC in the central nervous system.Protein kinase C (PKC),1 a family of phospholipid-dependent serine/threonine kinases and of which there are at least 12 subspecies, plays an important role in various cellular signal transductions (Nishizuka, 1984, 1988, 1992). Regardless of ubiquitous expression of PKCs in various tissues, the central nervous system abundantly contains several unique PKCs. In particular, the γ-subspecies of PKC (γ-PKC) is present only in the central nervous system and is thought to be involved in many neuronal functions including the formation of neural plasticity and memory (Nishizuka, 1986; Abeliovich et al., 1993a ,b; Tanaka and Nishizuka, 1994).PKC isozymes are divided into three subfamilies based on differences in the regulatory domain: conventional PKC (cPKC), novel PKC (nPKC), and atypical PKC (aPKC). Conventional PKCs have two common regions in the regulatory domain, C1 and C2. The C1 region has two cysteine-rich loops (zinc finger–like motifs) that interact with diacylglycerol (DG) or phorbol esters (Nishizuka, 1988; Ono et al., 1989). The C2 region mediates calcium binding (Ono et al., 1989) and is only present in cPKCs (Ono et al., 1988b ), although a region related to C2 region was recently reported in nPKC, a calcium-independent PKC (Parker and Dekker, 1997). Full activation of cPKCs, including γ-PKC, requires DG and calcium. The C1 region is also present in nPKC, and one of the cysteine-rich loops is found in aPKCs.Conventional PKCs and nPKCs, whose regulatory domains contain C1, are known to be translocated from the cytosol to particulate fraction when activated by DG or phorbol esters (Kraft et al., 1982). Therefore, the translocation of PKCs is a good marker of whether these enzymes are activated. Although this phenomenon is well known, the mechanism and physiological significance of PKC translocation have not yet been clarified. By conventional enzymological or immunohistochemical methods, it is impossible to observe the translocation of PKC in real time, in the same cells, and in living states, except in the investigation using fluorescent probes that directly bind PKC (Chen and Poenie, 1993). In addition, these fluorescent compounds are suggested to inhibit the activity of PKC itself at high concentration.To resolve these problems and to directly observe the translocation of γ-PKC in living cells, we produced a fusion protein of γ-PKC and green fluorescent protein (GFP). The GFP, isolated from jellyfish Aequorea victoria, has fluorescence without additional substrates and cofactors (Cubitt et al., 1995). Recent studies have revealed that GFP is a good candidate as a molecular reporter protein to monitor the alternation of protein localization, gene expression, and protein trafficking in living cells (Cubitt et al., 1995). In this study, we visualized and analyzed the translocation of γ-PKC–GFP fusion protein with confocal laser scanning fluorescence microscopy, using various stimulations, such as phorbol esters, Ca2+ ionophore, K+ depolarization, and receptor-mediated stimulus.  相似文献   

18.
Emerging evidence supports the ion channel mechanism for Alzheimer's disease pathophysiology wherein small β-amyloid (Aβ) oligomers insert into the cell membrane, forming toxic ion channels and destabilizing the cellular ionic homeostasis. Solid-state NMR-based data of amyloid oligomers in solution indicate that they consist of a double-layered β-sheets where each monomer folds into β-strand-turn-β-strand and the monomers are stacked atop each other. In the membrane, Aβ peptides are proposed to be β-type structures. Experimental structural data available from atomic force microscopy (AFM) imaging of Aβ oligomers in membranes reveal heterogeneous channel morphologies. Previously, we modeled the channels in a non-tilted organization, parallel with the cross-membrane normal. Here, we modeled a β-barrel-like organization. β-Barrels are common in transmembrane toxin pores, typically consisting of a monomeric chain forming a pore, organized in a single-layered β-sheet with antiparallel β-strands and a right-handed twist. Our explicit solvent molecular dynamics simulations of a range of channel sizes and polymorphic turns and comparisons of these with AFM image dimensions support a β-barrel channel organization. Different from the transmembrane β-barrels where the monomers are folded into a circular β-sheet with antiparallel β-strands stabilized by the connecting loops, these Aβ barrels consist of multimeric chains forming double β-sheets with parallel β-strands, where the strands of each monomer are connected by a turn. Although the Aβ barrels adopt the right-handed β-sheet twist, the barrels still break into heterogeneous, loosely attached subunits, in good agreement with AFM images and previous modeling. The subunits appear mobile, allowing unregulated, hence toxic, ion flux.  相似文献   

19.
20.
The aggregation behavior as a function of pH was studied for hydrolysates obtained by hydrolysis of soy protein isolate (SPI) and glycinin- and β-conglycinin-rich protein fractions with subtilisin Carlsberg. The substrates were hydrolyzed up to degrees of hydrolysis (DH) of 2.2% and 6.5%. Compared with nonhydrolyzed SPI, a decrease in solubility was observed for the hydrolysates of SPI [0.8% (w/v) protein, I = 0.03 M] around neutral pH. At pH 8.0, glycinin hydrolysates had a much lower solubility (∼43% and 60%, respectively, for DH 2.2% and 6.5%) than SPI and β-conglycinin-derived hydrolysates, which were almost completely soluble. Peptides that aggregated were all larger than 5 kDa, and as estimated by size-exclusion chromatography their composition was almost independent of the aggregation pH. The solubility of hydrolysates of SPIs with a varying glycinin and β-conglycinin composition showed that glycinin-derived peptides are the driving force for the lower solubility of SPI hydrolysates. The solubility of SPI hydrolysates at pH 8.0 was shown not to be the sum of that of glycinin and β-conglycinin hydrolysates. Assuming that the separate hydrolysis of glycinin and β-conglycinin did not differ from that in the mixture (SPI), this indicates that β-conglycinin-derived peptides have the ability to inhibit glycinin-derived peptide aggregation.  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