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 31 毫秒
1.
Dielectric spectroscopy of plant protoplasts   总被引:1,自引:0,他引:1       下载免费PDF全文
The relative permittivity and conductivity of the mesophyll protoplasts isolated from Brassica campestris leaves and Tulipa gesneriana petals were measured over a frequency range from 1kHz to 500 MHz.These protoplasts showed a broad dielectric dispersion, which was composed of three subdispersions, termed β1-, β2-, and β3-dispersion in increasing order of frequency.The three subdispersions were assigned to the Maxwell-Wagner dispersion caused by charging processes at the interfaces of the surface and internal membranes; the plasma membrane, the tonoplast, and the membranes of cytoplasmic organelles (e.g., chloroplasts, granules, etc) primarily contribute to the β1-, β2-, and β3-dispersion, respectively. The whole dielectric dispersion curve was satisfactorily interpreted in terms of a spherical cell model taking a large vacuole and cytoplasmic organelles into account. Using this model the capacitances of the plasma membranes and the tonoplasts were estimated to be 0.6-0.7 μF/cm2 and 0.9-1.0 μF/cm2, respectively.  相似文献   

2.
The dielectric behavior of a suspension of synchronized, spherical cells has been investigated in relation to the electrical parameters of certain cell structures. In the quasistatic approximation, Poisson's equations are solved for the respective diffusive media, and the local charge distributions are derived by taking into account the continuity equations. The results describe both α and β dispersion and reduce, in the corresponding limiting cases, to previous reports. The dependence of suspension permittivity in α-and β-dispersion ranges on the diffusive effects, the conductivity, and the permittivity of cytoplasm, of membrane, and of culture medium as well as on membrane thickness is pointed out. The possibility is pointed out of characterizing cellular behavior by means of the evolution of certain electrical and morphological parameters during cell cycle progression as well the effects of different stimuli on cellular systems derived by fast dielectric spectroscopy. © 1996 Wiley-Liss, Inc.  相似文献   

3.
Expression of human Bax, a cardinal regulator of mitochondrial membrane permeabilization, causes death in yeast. We screened a human cDNA library for suppressors of Bax-mediated yeast death and identified human 14-3-3β/α, a protein whose paralogs have numerous chaperone-like functions. Here, we show that, yeast cells expressing human 14-3-3β/α are able to complement deletion of the endogenous yeast 14-3-3 and confer resistance to a variety of different stresses including cadmium and cycloheximide. The expression of 14-3-3β/α also conferred resistance to death induced by the target of rapamycin inhibitor rapamycin and by starvation for the amino acid leucine, conditions that induce autophagy. Cell death in response to these autophagic stimuli was also observed in the macroautophagic-deficient atg1Δ and atg7Δ mutants. Furthermore, 14-3-3β/α retained its ability to protect against the autophagic stimuli in these autophagic-deficient mutants arguing against so called ‘autophagic death''. In line, analysis of cell death markers including the accumulation of reactive oxygen species, membrane integrity and cell surface exposure of phosphatidylserine indicated that 14-3-3β/α serves as a specific inhibitor of apoptosis. Finally, we demonstrate functional conservation of these phenotypes using the yeast homolog of 14-3-3: Bmh1. In sum, cell death in response to multiple stresses can be counteracted by 14-3-3 proteins.  相似文献   

4.
We report a systematic study of the condensation of plasmid DNA by oligocations with variation of the charge, Z, from +3 to +31. The oligocations include a series of synthetic linear ε-oligo(l-lysines), (denoted εKn, n = 3–10, 31; n is the number of lysines with the ligand charge Z = n+1) and branched α-substituted homologues of εK10: εYK10, εLK10 (Z = +11); εRK10, εYRK10 and εLYRK10 (Z = +21). Data were obtained by light scattering, UV absorption monitored precipitation assay and isothermal titration calorimetry in a wide range concentrations of DNA and monovalent salt (KCl, CKCl). The dependence of EC50 (ligand concentration at the midpoint of DNA condensation) on CKCl shows the existence of a salt-independent regime at low CKCl and a salt-dependent regime with a steep rise of EC50 with increase of CKCl. Increase of the ligand charge shifts the transition from the salt-independent to salt-dependent regime to higher CKCl. A novel and simple relationship describing the EC50 dependence on DNA concentration, charge of the ligand and the salt-dependent dissociation constant of the ligand–DNA complex is derived. For the ε-oligolysines εK6–εK10, the experimental dependencies of EC50 on CKCl and Z are well-described by an equation with a common set of parameters. Implications from our findings for understanding DNA condensation in chromatin are discussed.  相似文献   

5.
The flash-induced charge movements during the photocycle of light-adapted bacteriorhodopsin in purple membranes attached to a black lipid membrane were investigated under voltage clamp and current clamp conditions. Signal registration ranged from 200 ns to 30 s after flash excitation using a logarithmic clock, allowing the equally weighted measurement of the electrical phenomena over eight decades of time. The active pumping signals were separated from the passive system discharge on the basis of an equivalent circuit analysis. Both measuring methods were shown to yield equivalent results, but the charge translocation could be accurately monitored over the whole time range only under current clamp conditions. To describe the time course of the photovoltage signals a model based on distributed kinetics was found to be more appropriate than discrete first order processes suggesting the existence of conformational substates with distributed activation energies. The time course of the active charge displacement is characterised by a continuous relaxation time spectrum with three broad peaks plus an unresolved fast transient (<0.3 μs) of opposite polarity. The time constants and relative amplitudes (in brackets) derived from the peak rate constants and relative areas of the three bands are: τ1 = 32 μs (20%), τ2 = 0.89 ms (15%) and τ3 = 18 ms (65%) at 25°C in 150 mM KCl at pH7. The Arrhenius plots of the peak rate constants were linear yielding activation energies of EA1 = 57 kJ/mol, EA2 = 52 kJ/mol, and EA3 = 44 kJ/mol. The electrical signal at 890 μs has no counterpart in the photocycle of bacteriorhodopsin suspensions. Fits with a sum of exponentials required 5 to 6 components and were not reproducible. Analysis of photoelectrical signals with continuous relaxation time spectra gave equally good fits with fewer parameters and were well reproducible.  相似文献   

6.
7.
Despite its essential role in the yeast cell wall, the exact composition of the β-(1,6)-glucan component is not well characterized. While solubilizing the cell wall alkali-insoluble fraction from a wild type strain of Saccharomyces cerevisiae using a recombinant β-(1,3)-glucanase followed by chromatographic characterization of the digest on an anion exchange column, we observed a soluble polymer that eluted at the end of the solvent gradient run. Further characterization indicated this soluble polymer to have a molecular mass of ∼38 kDa and could be hydrolyzed only by β-(1,6)-glucanase. Gas chromatographymass spectrometry and NMR (1H and 13C) analyses confirmed it to be a β-(1,6)-glucan polymer with, on average, branching at every fifth residue with one or two β-(1,3)-linked glucose units in the side chain. This polymer peak was significantly reduced in the corresponding digests from mutants of the kre genes (kre9 and kre5) that are known to play a crucial role in the β-(1,6)-glucan biosynthesis. In the current study, we have developed a biochemical assay wherein incubation of UDP-[14C]glucose with permeabilized S. cerevisiae yeasts resulted in the synthesis of a polymer chemically identical to the branched β-(1,6)-glucan isolated from the cell wall. Using this assay, parameters essential for β-(1,6)-glucan synthetic activity were defined.The cell wall of Saccharomyces cerevisiae and other yeasts contains two types of β-glucans. In the former yeast, branched β-(1,3)-glucan accounts for ∼50–55%, whereas β-(1,6)-glucan represents 10–15% of the total yeast cell wall polysaccharides, each chain of the latter extending up to 140–350 glucose residues in length. The amount of 3,6-branched glucose residues varies with the yeast species: 7, 15, and 75% in S. cerevisiae, Candida albicans, and Schizosaccharomyces pombe, respectively (1). β-(1,6)-Glucan stabilizes the cell wall, since it plays a central role as a linker for specific cell wall components, including β-(1,3)-glucan, chitin, and mannoproteins (2, 3). However, the exact structure of the β-(1,6)-glucan and the mode of biosynthesis of this polymer are largely unknown. In S. pombe, immunodetection studies suggested that synthesis of this polymer backbone begins in the endoplasmic reticulum, with extension occurring in the Golgi (4) and final processing at the plasma membrane. In S. cerevisiae, Montijn and co-workers (5), by immunogold labeling, detected β-(1,6)-glucan at the plasma membrane, suggesting that the synthesis takes place largely at the cell surface.More than 20 genes, including the KRE gene family (14 members) and their homologues, SKN1 and KNH1, have been reported to be involved in β-(1,6)-glucan synthesis in S. cerevisiae, C. albicans, and Candida glabrata (610). Among all of these genes, the ones that seem to play the major synthetic role are KRE5 and KRE9, since their disruption caused significant reduction (100 and 80%, respectively, relative to wild type) in the cell wall β-(1,6)-glucan content (1113).To date, the biochemical reaction responsible for the synthesis of β-(1,6)-glucan and the product synthesized remained unknown. Indeed, in most cases, when membrane preparations are incubated with UDP-glucose, only linear β-(1,3)-glucan polymers are produced, although some studies have reported the production of low amounts of β-(1,6)-glucans by membrane preparations (1417). These data suggest that disruption of the fungal cell prevents or at least has a strong negative effect on β-(1,6)-glucan synthesis. The use of permeabilized cells, which allows substrates, such as nucleotide sugar precursors, to be readily transported across the plasma membrane, is an alternative method to study in situ cell wall enzyme activities (1822). A number of methods have been developed to permeabilize the yeast cell wall (23), of which osmotic shock was successfully used to demonstrate β-(1,3)-glucan and chitin synthase activities (20, 24). Herein, we describe the biochemical activity responsible for β-(1,6)-glucan synthesis using permeabilized S. cerevisiae cells and UDP-[14C]glucose as a substrate. We also have analyzed the physicochemical parameters of this activity and chemically characterized the end product and its structural organization within the mature yeast cell wall.  相似文献   

8.
MHC class II molecules are composed of one α-chain and one β-chain whose membrane distal interface forms the peptide binding groove. Most of the existing knowledge on MHC class II molecules comes from the cis-encoded variants where the α- and β-chain are encoded on the same chromosome. However, trans-encoded class II MHC molecules, where the α- and β-chain are encoded on opposite chromosomes, can also be expressed. We have studied the trans-encoded class II HLA molecule DQ2.3 (DQA1*03:01/DQB1*02:01) that has received particular attention as it may explain the increased risk of certain individuals to type 1 diabetes. We report the x-ray crystal structure of this HLA molecule complexed with a gluten epitope at 3.05 Å resolution. The gluten epitope, which is the only known HLA-DQ2.3-restricted epitope, is preferentially recognized in the context of the DQ2.3 molecule by T-cell clones of a DQ8/DQ2.5 heterozygous celiac disease patient. This preferential recognition can be explained by improved HLA binding as the epitope combines the peptide-binding motif of DQ2.5 (negative charge at P4) and DQ8 (negative charge at P1). The analysis of the structure of DQ2.3 together with all other available DQ crystal structures and sequences led us to categorize DQA1 and DQB1 genes into two groups where any α-chain and β-chain belonging to the same group are expected to form a stable heterodimer.  相似文献   

9.
10.
Plants evoke innate immunity against microbial challenges upon recognition of pathogen-associated molecular patterns (PAMPs), such as fungal cell wall chitin. Nevertheless, pathogens may circumvent the host PAMP-triggered immunity. We previously reported that the ascomycete Magnaporthe oryzae, a famine-causing rice pathogen, masks cell wall surfaces with α-1,3-glucan during invasion. Here, we show that the surface α-1,3-glucan is indispensable for the successful infection of the fungus by interfering with the plant''s defense mechanisms. The α-1,3-glucan synthase gene MgAGS1 was not essential for infectious structure development but was required for infection in M. oryzae. Lack or degradation of surface α-1,3-glucan increased fungal susceptibility towards chitinase, suggesting the protective role of α-1,3-glucan against plants'' antifungal enzymes during infection. Furthermore, rice plants secreting bacterial α-1,3-glucanase (AGL-rice) showed strong resistance not only to M. oryzae but also to the phylogenetically distant ascomycete Cochlioborus miyabeanus and the polyphagous basidiomycete Rhizoctonia solani; the histocytochemical analysis of the latter two revealed that α-1,3-glucan also concealed cell wall chitin in an infection-specific manner. Treatment with α-1,3-glucanase in vitro caused fragmentation of infectious hyphae in R. solani but not in M. oryzae or C. miyabeanus, indicating that α-1,3-glucan is also involved in maintaining infectious structures in some fungi. Importantly, rapid defense responses were evoked (a few hours after inoculation) in the AGL-rice inoculated with M. oryzae, C. miyabeanus and R. solani as well as in non-transgenic rice inoculated with the ags1 mutant. Taken together, our results suggest that α-1,3-glucan protected the fungal cell wall from degradative enzymes secreted by plants even from the pre-penetration stage and interfered with the release of PAMPs to delay innate immune defense responses. Because α-1,3-glucan is nondegradable in plants, it is reasonable that many fungal plant pathogens utilize α-1,3-glucan in the innate immune evasion mechanism and some in maintaining the structures.  相似文献   

11.
12.
Biosynthesis of the mycobacterial cell wall relies on the activities of many enzymes, including several glycosyltransferases (GTs). The polymerizing galactofuranosyltransferase GlfT2 (Rv3808c) synthesizes the bulk of the galactan portion of the mycolyl-arabinogalactan complex, which is the largest component of the mycobacterial cell wall. We used x-ray crystallography to determine the 2.45-Å resolution crystal structure of GlfT2, revealing an unprecedented multidomain structure in which an N-terminal β-barrel domain and two primarily α-helical C-terminal domains flank a central GT-A domain. The kidney-shaped protomers assemble into a C4-symmetric homotetramer with an open central core and a surface containing exposed hydrophobic and positively charged residues likely involved with membrane binding. The structure of a 3.1-Å resolution complex of GlfT2 with UDP reveals a distinctive mode of nucleotide recognition. In addition, models for the binding of UDP-galactofuranose and acceptor substrates in combination with site-directed mutagenesis and kinetic studies suggest a mechanism that explains the unique ability of GlfT2 to generate alternating β-(1→5) and β-(1→6) glycosidic linkages using a single active site. The topology imposed by docking a tetrameric assembly onto a membrane bilayer also provides novel insights into aspects of processivity and chain length regulation in this and possibly other polymerizing GTs.  相似文献   

13.
The paralyzed zebrafish strain relaxed carries a null mutation for the skeletal muscle dihydropyridine receptor (DHPR) β1a subunit. Lack of β1a results in (i) reduced membrane expression of the pore forming DHPR α1S subunit, (ii) elimination of α1S charge movement, and (iii) impediment of arrangement of the DHPRs in groups of four (tetrads) opposing the ryanodine receptor (RyR1), a structural prerequisite for skeletal muscle-type excitation-contraction (EC) coupling. In this study we used relaxed larvae and isolated myotubes as expression systems to discriminate specific functions of β1a from rather general functions of β isoforms. Zebrafish and mammalian β1a subunits quantitatively restored α1S triad targeting and charge movement as well as intracellular Ca2+ release, allowed arrangement of DHPRs in tetrads, and most strikingly recovered a fully motile phenotype in relaxed larvae. Interestingly, the cardiac/neuronal β2a as the phylogenetically closest, and the ancestral housefly βM as the most distant isoform to β1a also completely recovered α1S triad expression and charge movement. However, both revealed drastically impaired intracellular Ca2+ transients and very limited tetrad formation compared with β1a. Consequently, larval motility was either only partially restored (β2a-injected larvae) or not restored at all (βM). Thus, our results indicate that triad expression and facilitation of 1,4-dihydropyridine receptor (DHPR) charge movement are common features of all tested β subunits, whereas the efficient arrangement of DHPRs in tetrads and thus intact DHPR-RyR1 coupling is only promoted by the β1a isoform. Consequently, we postulate a model that presents β1a as an allosteric modifier of α1S conformation enabling skeletal muscle-type EC coupling.Excitation-contraction (EC)3 coupling in skeletal muscle is critically dependent on the close interaction of two distinct Ca2+ channels. Membrane depolarizations of the myotube are sensed by the voltage-dependent 1,4-dihydropyridine receptor (DHPR) in the sarcolemma, leading to a rearrangement of charged amino acids (charge movement) in the transmembrane segments S4 of the pore-forming DHPR α1S subunit (1, 2). This conformational change induces via protein-protein interaction (3, 4) the opening of the sarcoplasmic type-1 ryanodine receptor (RyR1) without need of Ca2+ influx through the DHPR (5). The release of Ca2+ from the sarcoplasmic reticulum via RyR1 consequently induces muscle contraction. The protein-protein interaction mechanism between DHPR and RyR1 requires correct ultrastructural targeting of both channels. In Ca2+ release units (triads and peripheral couplings) of the skeletal muscle, groups of four DHPRs (tetrads) are coupled to every other RyR1 and hence are geometrically arranged following the RyR-specific orthogonal arrays (6).The skeletal muscle DHPR is a heteromultimeric protein complex, composed of the voltage-sensing and pore-forming α1S subunit and auxiliary subunits β1a, α2δ-1, and γ1 (7). While gene knock-out of the DHPR γ1 subunit (8, 9) and small interfering RNA knockdown of the DHPR α2δ-1 subunit (10-12) have indicated that neither subunit is essential for coupling of the DHPR with RyR1, the lack of the α1S or of the intracellular β1a subunit is incompatible with EC coupling and accordingly null model mice die perinatally due to asphyxia (13, 14). β subunits of voltage-gated Ca2+ channels were repeatedly shown to be responsible for the facilitation of α1 membrane insertion and to be potent modulators of α1 current kinetics and voltage dependence (15, 16). Whether the loss of EC coupling in β1-null mice was caused by decreased DHPR membrane expression or by the lack of a putative specific contribution of the β subunit to the skeletal muscle EC coupling apparatus (17, 18) was not clearly resolved. Recently, other β-functions were identified in skeletal muscle using the β1-null mutant zebrafish relaxed (19, 20). Like the β1-knock-out mouse (14) zebrafish relaxed is characterized by complete paralysis of skeletal muscle (21, 22). While β1-knock-out mouse pups die immediately after birth due to respiratory paralysis (14), larvae of relaxed are able to survive for several days because of oxygen and metabolite diffusion via the skin (23). Using highly differentiated myotubes that are easy to isolate from these larvae, the lack of EC coupling could be described by quantitative immunocytochemistry as a moderate ∼50% reduction of α1S membrane expression although α1S charge movement was nearly absent, and, most strikingly, as the complete lack of the arrangement of DHPRs in tetrads (19). Thus, in skeletal muscle the β subunit enables EC coupling by (i) enhancing α1S membrane targeting, (ii) facilitating α1S charge movement, and (iii) enabling the ultrastructural arrangement of DHPRs in tetrads.The question arises, which of these functions are specific for the skeletal muscle β1a and which ones are rather general properties of Ca2+ channel β subunits. Previous reconstitution studies made in the β1-null mouse system (24, 25) using different β subunit constructs (26) did not allow differentiation between β-induced enhancement of non-functional α1S membrane expression and the facilitation of α1S charge movement, due to the lack of information on α1S triad expression levels. Furthermore, the β-induced arrangement of DHPRs in tetrads was not detected as no ultrastructural information was obtained.In the present study, we established zebrafish mutant relaxed as an expression system to test different β subunits for their ability to restore skeletal muscle EC coupling. Using isolated myotubes for in vitro experiments (19, 27) and complete larvae for in vivo expression studies (28-31) and freeze-fracture electron microscopy, a clear differentiation between the major functional roles of β subunits was feasible in the zebrafish system. The cloned zebrafish β1a and a mammalian (rabbit) β1a were shown to completely restore all parameters of EC coupling when expressed in relaxed myotubes and larvae. However, the phylogenetically closest β subunit to β1a, the cardiac/neuronal isoform β2a from rat, as well as the ancestral βM isoform from the housefly (Musca domestica), could recover functional α1S membrane insertion, but led to very restricted tetrad formation when compared with β1a, and thus to impaired DHPR-RyR1 coupling. This impairment caused drastic changes in skeletal muscle function.The present study shows that the enhancement of functional α1S membrane expression is a common function of all the tested β subunits, from β1a to even the most distant βM, whereas the effective formation of tetrads and thus proper skeletal muscle EC coupling is an exclusive function of the skeletal muscle β1a subunit. In context with previous studies, our results suggest a model according to which β1a acts as an allosteric modifier of α1S conformation. Only in the presence of β1a, the α1S subunit is properly folded to allow RyR1 anchoring and thus skeletal muscle-type EC coupling.  相似文献   

14.
The L-type calcium channel (LTCC) has a variety of physiological roles that are critical for the proper function of many cell types and organs. Recently, a member of the zinc-regulating family of proteins, ZnT-1, was recognized as an endogenous inhibitor of the LTCC, but its mechanism of action has not been elucidated. In the present study, using two-electrode voltage clamp recordings in Xenopus oocytes, we demonstrate that ZnT-1-mediated inhibition of the LTCC critically depends on the presence of the LTCC regulatory β-subunit. Moreover, the ZnT-1-induced inhibition of the LTCC current is also abolished by excess levels of the β-subunit. An interaction between ZnT-1 and the β-subunit, as demonstrated by co-immunoprecipitation and by fluorescence resonance energy transfer, is consistent with this result. Using surface biotinylation and total internal reflection fluorescence microscopy in HEK293 cells, we show a ZnT-1-dependent decrease in the surface expression of the pore-forming α1-subunit of the LTCC. Similarly, a decrease in the surface expression of the α1-subunit is observed following up-regulation of the expression of endogenous ZnT-1 in rapidly paced cultured cardiomyocytes. We conclude that ZnT-1-mediated inhibition of the LTCC is mediated through a functional interaction of ZnT-1 with the LTCC β-subunit and that it involves a decrease in the trafficking of the LTCC α1-subunit to the surface membrane.  相似文献   

15.
Cationic antimicrobial peptides/proteins (AMPs) are important components of the host innate defense mechanisms against invading microorganisms. Here we demonstrate that OprI (outer membrane protein I) of Pseudomonas aeruginosa is responsible for its susceptibility to human ribonuclease 7 (hRNase 7) and α-helical cationic AMPs, instead of surface lipopolysaccharide, which is the initial binding site of cationic AMPs. The antimicrobial activities of hRNase 7 and α-helical cationic AMPs against P. aeruginosa were inhibited by the addition of exogenous OprI or anti-OprI antibody. On modification and internalization of OprI by hRNase 7 into cytosol, the bacterial membrane became permeable to metabolites. The lipoprotein was predicted to consist of an extended loop at the N terminus for hRNase 7/lipopolysaccharide binding, a trimeric α-helix, and a lysine residue at the C terminus for cell wall anchoring. Our findings highlight a novel mechanism of antimicrobial activity and document a previously unexplored target of α-helical cationic AMPs, which may be used for screening drugs to treat antibiotic-resistant bacterial infection.  相似文献   

16.
17.
18.

Introduction

TNFα is a proinflammatory cytokine that plays a central role in the pathogenesis of rheumatoid arthritis (RA). We investigated the effects of certolizumab pegol, a TNFα blocker, on endothelial cell function and angiogenesis.

Methods

Human dermal microvascular endothelial cells (HMVECs) were stimulated with TNFα with or without certolizumab pegol. TNFα-induced adhesion molecule expression and angiogenic chemokine secretion were measured by cell surface ELISA and angiogenic chemokine ELISA, respectively. We also examined the effect of certolizumab pegol on TNFα-induced myeloid human promyelocytic leukemia (HL-60) cell adhesion to HMVECs, as well as blood vessels in RA synovial tissue using the Stamper-Woodruff assay. Lastly, we performed HMVEC chemotaxis, and tube formation.

Results

Certolizumab pegol significantly blocked TNFα-induced HMVEC cell surface angiogenic E-selectin, vascular cell adhesion molecule-1 and intercellular adhesion molecule-1 expression and angiogenic chemokine secretion (P < 0.05). We found that certolizumab pegol significantly inhibited TNFα-induced HL-60 cell adhesion to HMVECs (P < 0.05), and blocked HL-60 cell adhesion to RA synovial tissue vasculature (P < 0.05). TNFα also enhanced HMVEC chemotaxis compared with the negative control group (P < 0.05) and this chemotactic response was significantly reduced by certolizumab pegol (P < 0.05). Certolizumab pegol inhibited TNFα-induced HMVEC tube formation on Matrigel (P < 0.05).

Conclusion

Our data support the hypothesis that certolizumab pegol inhibits TNFα-dependent leukocyte adhesion and angiogenesis, probably via inhibition of angiogenic adhesion molecule expression and angiogenic chemokine secretion.  相似文献   

19.
N-Glycosylation of integrin α5β1 plays a crucial role in cell spreading, cell migration, ligand binding, and dimer formation, but the detailed mechanisms by which N-glycosylation mediates these functions remain unclear. In a previous study, we showed that three potential N-glycosylation sites (α5S3–5) on the β-propeller of the α5 subunit are essential to the functional expression of the subunit. In particular, site 5 (α5S5) is the most important for its expression on the cell surface. In this study, the function of the N-glycans on the integrin β1 subunit was investigated using sequential site-directed mutagenesis to remove the combined putative N-glycosylation sites. Removal of the N-glycosylation sites on the I-like domain of the β1 subunit (i.e. the Δ4-6 mutant) decreased both the level of expression and heterodimeric formation, resulting in inhibition of cell spreading. Interestingly, cell spreading was observed only when the β1 subunit possessed these three N-glycosylation sites (i.e. the S4-6 mutant). Furthermore, the S4-6 mutant could form heterodimers with either α5S3-5 or α5S5 mutant of the α5 subunit. Taken together, the results of the present study reveal for the first time that N-glycosylation of the I-like domain of the β1 subunit is essential to both the heterodimer formation and biological function of the subunit. Moreover, because the α5S3-5/β1S4-6 mutant represents the minimal N-glycosylation required for functional expression of the β1 subunit, it might also be useful for the study of molecular structures.Integrin is a heterodimeric glycoprotein that consists of both an α and a β subunit (1). The interaction between integrin and the extracellular matrix is essential to both physiologic and pathologic events, such as cell migration, development, cell viability, immune homeostasis, and tumorigenesis (2, 3). Among the integrin superfamily, β1 integrin can combine with 12 distinct α subunits (α1–11, αv) to form heterodimers, thereby acquiring a wide variety of ligand specificity (1, 4). Integrins are thought to be regulated by inside-out signaling mechanisms that provoke conformational changes, which modulate the affinity of integrin for the ligand (5). However, an increasing body of evidence suggests that cell-surface carbohydrates mediate a variety of interactions between integrin and its extracellular environment, thereby affecting integrin activity and possibly tumor metastasis as well (68).Guo et al. (9) reported that an increase in β1–6-GlcNAc sugar chains on the integrin β1 subunit stimulated cell migration. In addition, elevated sialylation of the β1 subunit, because of Ras-induced STGal-I transferase activity, also induced cell migration (10, 11). Conversely, cell migration and spreading were reduced by the addition of a bisecting GlcNAc, which is a product of N-acetylglucosaminyltransferase III (GnT-III),2 to the α5β1 and α3β1 integrins (12, 13). Alterations of N-glycans on integrins might also regulate their cis interactions with membrane-associated proteins, including the epidermal growth factor receptor, the galectin family, and the tetraspanin family of proteins (1419).In addition to the positive and negative regulatory effects of N-glycan, several research groups have reported that N-glycans must be present on integrin α5β1 for the αβ heterodimer formation and proper integrin-matrix interactions. Consistent with this hypothesis, in the presence of the glycosylation inhibitor, tunicamycin, normal integrin-substrate binding and transport to the cell surface are inhibited (20). Moreover, treatment of purified integrin with N-glycosidase F blocked both the inherent association of the subunits and the interaction between integrin and fibronectin (FN) (21). These results suggest that N-glycosylation is essential to the functional expression of α5β1. However, because integrin α5β1 contains 26 potential N-linked glycosylation sites, 14 in the α subunit and 12 in the β subunit, identification of the sites that are essential to its biological functions is key to understanding the molecular mechanisms by which N-glycans alter integrin function. Recently, our group determined that N-glycosylation of the β-propeller domain on the α5 subunit is essential to both heterodimerization and biological functions of the subunit. Furthermore, we determined that sites 3–5 are the most important sites for α5 subunit-mediated cell spreading and migration on FN (22). The purpose of this study was to clarify the roles of N-glycosylation of the β1 subunit. Therefore, we performed combined substitutions in the putative N-glycosylation sites by replacement of asparagine residues with glutamine residues. We subsequently introduced these mutated genes into β1-deficient epithelial cells (GE11). The results of these mutation experiments revealed that the N-glycosylation sites on the I-like domain of the β1 subunit, sites number 4–6 (S4-6), are essential to both heterodimer formation and biological functions, such as cell spreading.  相似文献   

20.
Endo T  Kawano S  Yamano K 《EMBO reports》2011,12(2):94-95
A study recently published in EMBO reports solves the solution structure of E. coli BamE, thus providing the basis for a better understanding of the mechanism of β-barrel assembly in bacterial and mitochondrial outer membranes.EMBO Rep (2011) advance online publication. doi: 10.1038/embor.2010.202β-barrel membrane proteins are found exclusively in the outer membrane of Gram-negative bacteria and the outer membranes of eukaryotic organelles of prokaryotic origin, mitochondria and chloroplasts. In contrast to the inner membrane, the bacterial outer membrane is an asymmetrical bilayer that consists mainly of lipopolysaccharides in the outer leaflet and phospholipids in the inner leaflet. Bacterial β-barrel outer membrane proteins (OMPs) mediate many cellular functions, for example, passive or selective diffusion of small molecules through the β-barrel pores across the outer membrane. By contrast, only a few mitochondrial β-barrel outer membrane proteins (MBOMPs) have been identified so far. The central machineries that mediate insertion and assembly of OMPs/MBOMPs are the β-barrel assembly machine (BAM) complex in the bacterial outer membrane and the topogenesis of outer-membrane β-barrel proteins (TOB)/sorting and assembly machinery (SAM) complex in the mitochondrial outer membrane (Knowles et al, 2009; Endo & Yamano, 2010; Stroud et al, 2010; Fig 1). However, the molecular mechanisms of β-barrel protein topogenesis in bacterial and mitochondrial outer membranes remain poorly understood.Open in a separate windowFigure 1β-barrel protein assembly in bacterial and mitochondrial outer membranes. (A) Bacteria. Ribbon models of the structures of the Sec complex, SurA, BamA (Clantin et al, 2007; Kim et al, 2007), BamE and OMP. The upper and lower inserts show the surface of BamE (residues 20–108; viewed after approximately 90° rotation of the ribbon model around the horizontal axis toward the reader). Residues important for BamD binding are shown in red and residues with NMR signals that were perturbed by BamD binding are shown in yellow. The residue (Phe 74) important for PG binding is shown in red and the residues with NMR signals that were perturbed by PG binding are shown in yellow. (B) Mitochondria. Ribbon models were drawn for the structures of small Tim and MBOMP. IM, inner membrane; IMS, intermembrane space; MBOMP, mitochondrial β-barrel outer membrane protein; OM, outer membrane; OMP, outer membrane protein; PG, phosphatidylglycerol; POTRA, polypeptide transport-associated domain.Bacterial OMPs are synthesized in the cytosol as precursor proteins with an amino-terminal signal sequence that guides the proteins to the Sec machinery for crossing the inner membrane and is cleaved off in the periplasm. Periplasmic chaperones then escort OMPs through the aqueous periplasmic space in a partly unfolded state. On reaching the outer membrane, OMPs assemble into a β-barrel structure and insert into the outer membrane with the help of the BAM complex. The bacterial OMP insertion pathway can be compared to the assembly pathway of MBOMPs from the mitochondrial intermembrane space into the outer membrane. MBOMPs are synthesized in the cytosol and imported into the intermembrane space by the outer membrane translocator TOM40. The subsequent chaperone-mediated escort across the intermembrane space and insertion into the outer membrane by the TOB complex is similar to the OMP assembly process. Notably, the BAM and TOB complexes share the homologous β-barrel proteins BamA and Tob55/Sam50, respectively, as the central components of their insertion machineries. The BAM complex in Escherichia coli consists of BamA (YaeT/Omp85) and four accessory lipoproteins: BamB (YfgL), BamC (NlpB), BamD (YfiO) and BamE (SmpA). BamA and BamD are essential for cell growth, yet deletion of dispensable BamB, BamC or BamE leads to outer membrane defects manifested in hypersensitivity to antibiotics. Although BamAB and BamCDE can form distinct subcomplexes, they become functional only after formation of the entire BAM complex with all five subunits (Hagan et al, 2010).In this issue of EMBO reports, Knowles et al (2011) solve the nuclear magnetic resonance (NMR) solution structure of E. coli BamE, which sheds light on the roles of one of the Bam subunits in β-barrel protein assembly. The structure of BamE consists of a three-stranded antiparallel β-sheet packed against a pair of α-helices (Fig 1).As the ΔbamE mutant cannot grow in the presence of vancomycin, the authors identify functionally important residues of BamE by testing the effects of amino-acid substitutions in BamE on its inability to complement the growth defects of ΔbamE, without destabilizing BamE itself. Many of the identified residues are conserved among BamE proteins from different organisms and map to a single surface area on BamE. Interestingly, NMR signals of the residues around this region are sensitive to the addition of micelles containing the lipid phosphatidylglycerol, but not phosphatidylethanolamine or cardiolipin. In parallel, the authors analyse perturbation of the NMR spectra of BamE after the addition of purified BamB, C and D proteins. Only BamD affects the NMR spectra of BamE, and the BamD interacting region of BamE is found to overlap partly with the residues involved in phosphatidylglycerol binding. As the addition of BamD and phosphatidylglycerol have different effects on the NMR spectra of BamE, the binding of BamD and phosphatidylglycerol to BamE seem to take place simultaneously. What is the biological relevance of the observed interactions of BamE with both BamD and phosphatidylglycerol? As phosphatidylglycerol was found to help the insertion of OMPs into lipid liposomes (Patel et al, 2009), BamE might recruit the BAM complex through BamD to phosphatidylglycerol-rich regions in the outer membrane, or might directly recruit phosphatidylglycerol to form assembly points for OMP insertion and folding.What are the roles of other subunits of the BAM complex in β-barrel protein assembly? The essential subunit of the E. coli BAM complex BamA consists of two domains: the N-terminal polypeptide transport-associated (POTRA) domain repeat in the periplasm and the carboxy-terminal β-barrel domain, embedded in the outer membrane. The number of POTRA domains ranges from one to five in BamA homologues from different organisms. Of these POTRA domains, the one nearest to the C-terminal that is most connected to the β-barrel domain is essential for cell viability and its deletion leads to disassembly of the BAM complex (Kim et al, 2007). Structural studies of the E. coli BamA POTRA domains suggest that each POTRA domain has a common fold, whereas conformational rigidity might differ between inter-domain linkers (Gatzeva-Topalova et al, 2010; Fig 1). As individual POTRA domains have some affinity for unfolded substrate proteins, the periplasmic tandem POTRA repeat probably provides several substrate binding sites that slide the substrate progressively towards the BamA β-barrel domain. The β-barrel domain of BamA probably functions as a scaffold to facilitate the formation of β-strands, possibly through β-augmentation and subsequent spontaneous membrane insertion of the β-barrel. Yet, it is not clear whether this cradle for β-strand formation is provided by the pore formed within the monomer or oligomeric forms of the BamA β-barrel domain. Alternatively, membrane insertion and folding of OMPs might take place at the interface between BamA and the outer membrane lipid bilayer.How much of the β-barrel assembly process is conserved during the evolution of mitochondria from Gram-negative bacteria? Although the central subunits BamA and Tob55 of the BAM and TOB complexes are conserved, other subunits of these complexes are unrelated to each other. The POTRA domains of BamA are essential for recognition and assembly of bacterial OMPs, whereas that of Tob55 is dispensable for MBOMP assembly in the mitochondrial outer membrane. Nevertheless, the mitochondrial TOB complex facilitates assembly of bacterial OMPs at low efficiency (Walther et al, 2009) and, in turn, the bacterial BAM complex can mediate assembly of mitochondrial porin. Therefore, the basic mechanism of β-barrel assembly in the outer membranes of bacteria and mitochondria seems to be conserved. High-resolution structures of each component of the BAM and TOB complexes—including that of BamE in this study—will thus provide the basis for a better understanding of the mechanism of β-barrel assembly in evolutionarily related bacterial and mitochondrial outer membranes.  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