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 921 毫秒
1.
Alzheimer's disease (AD) is characterized by large numbers of senile plaques in the brain that consist of fibrillar aggregates of 40- and 42-residue amyloid-beta (Abeta) peptides. However, the degree of dementia in AD correlates better with the concentration of soluble Abeta species assayed biochemically than with histologically determined plaque counts, and several investigators now propose that soluble aggregates of Abeta are the neurotoxic agents that cause memory deficits and neuronal loss. These endogenous aggregates are minor components in brain extracts from AD patients and transgenic mice that express human Abeta, but several species have been detected by gel electrophoresis in sodium dodecylsulfate (SDS) and isolated by size exclusion chromatography (SEC). Endogenous Abeta aggregation is stimulated at cellular interfaces rich in lipid rafts, and anionic micelles that promote Abeta aggregation in vitro may be good models of these interfaces. We previously found that micelles formed in dilute SDS (2 mM) promote Abeta(1-40) fiber formation by supporting peptide interaction on the surface of a single micelle complex. In contrast, here we report that monomeric Abeta(1-42) undergoes an immediate conversion to a predominant beta-structured conformation in 2 mM SDS which does not proceed to amyloid fibrils. The conformational change is instead rapidly followed by the near quantitative conversion of the 4 kDa monomer SDS gel band to 8-14 kDa bands consistent with dimers through tetramers. Removal of SDS by dialysis gave a shift in the predominant SDS gel bands to 30-60 kDa. While these oligomers resemble the endogenous aggregates, they are less stable. In particular, they do not elute as discrete species on SEC, and they are completed disaggregated by boiling in 1% SDS. It appears that endogenous oligomeric Abeta aggregates are stabilized by undefined processes that have not yet been incorporated into in vitro Abeta aggregation procedures.  相似文献   

2.
2-Cys peroxiredoxins (2-Cys Prxs) are ubiquitous peroxidases with important roles in cellular antioxidant defense and hydrogen peroxide-mediated signaling. Post-translational modifications of conserved cysteines cause the transition from low to high molecular weight oligomers, triggering the functional change from peroxidase to molecular chaperone. However, it remains unclear how non-covalent interactions of 2-Cys Prx with metabolites modulate the quaternary structure. Here, we disclose that ATP and Mg(2+) (ATP/Mg) promote the self-polymerization of chloroplast 2-Cys Prx (polypeptide 23.5 kDa) into soluble higher order assemblies (>2 MDa) that proceed to insoluble aggregates beyond 5 mM ATP. Remarkably, the withdrawal of ATP or Mg(2+) brings soluble oligomers and insoluble aggregates back to the native conformation without compromising the associated functions. As confirmed by transmission electron microscopy, ATP/Mg drive the toroid-like decamers (diameter 13 nm) to the formation of large sphere-like particles (diameter ~30 nm). Circular dichroism studies on ATP-labeled 2-Cys Prx reveal that ATP/Mg enhance the proportion of β-sheets with the concurrent decrease in the content of α-helices. In line with this observation, the formation of insoluble aggregates is strongly prevented by 2,2,2-trifluoroethanol, a cosolvent employed to induce α-helical conformations. We further find that the response of self-polymerization to ATP/Mg departs abruptly from that of the associated peroxidase and chaperone activities when two highly conserved residues, Arg(129) and Arg(152), are mutated. Collectively, our data uncover that non-covalent interactions of ATP/Mg with 2-Cys Prx modulate dynamically the quaternary structure, thereby coupling the non-redox chemistry of cell energy with redox transformations at cysteine residues.  相似文献   

3.
A subunit with carboxymethyl cellulase (CMCase) activity was isolated from the cellulosomes of Clostridium thermocellum after dissociation of the cellulosomes by a mild sodium dodecyl sulfate (SDS) treatment. The subunit displayed only one protein band of 51 kDa on SDS-polyacrylamide gel electrophoresis (SDS-PAGE), but after boiling with SDS it had 3 bands of 60, 56, and 48 kDa. Prolonged incubation with SDS changed the subunit to display exclusively the 48-kDa band after boiling. The 51-kDa subunit was presumably a partially denatured form, and differentiated into 3 species with apparent M(r) of 60, 56, and 48 k through deglycosylation in SDS solution. Enzymatic properties of the 51-kDa subunit resembled those of the endoglucanase A which was purified from the culture fluid and from a E. coli clone with exceptions of temperature and pH optima.  相似文献   

4.
DNA binding properties of a 110 kDa nucleolar protein.   总被引:8,自引:2,他引:6       下载免费PDF全文
M Sapp  R Knippers    A Richter 《Nucleic acids research》1986,14(17):6803-6820
A single strand specific DNA binding protein was purified to homogeneity from calf thymus nucleoprotein. The monomeric protein is elongated in shape and has a molecular mass of 110 kDa. Since immunocytochemistry revealed that the protein is predominantly located in the nucleolus we refer to it as the 110 kDa nucleolar protein. The protein binds not only to single stranded DNA but also to single stranded RNA, including homopolymeric synthetic RNA. We have used the single stranded DNA binding properties of the 110 kDa protein in model studies to investigate its effects on the configuration of nucleic acid. Our results are: only 50-55 protein molecules are sufficient to saturate all binding sites on the 6408 nucleotides of phage fd DNA; protein binding cause a compaction of single stranded DNA; large nucleoprotein aggregates are formed in the presence of divalent cations; this is due to protein-protein interactions which occur at moderately high concentrations of magnesium-, calcium or manganese ions; the protein induces the reassociation of complementary nucleic acid sequences. We speculate that the 110 kDa protein performs similar reactions in vivo and may have a function related to the processing and packaging of preribosomal RNA.  相似文献   

5.
Outer membrane protein of Escherichia coli prepared for polyacrylamide gel electrophoresis by solubilization of the membrane in an organic solvent followed by dialysis into sodium dodecyl sulfate (SDS) solution or by solublization of the membrane directly in SDS solution followed by dialysis into a SDS-urea solution and brief heating at 100 °C resulted in a simple polypeptide profile on SDS-containing gels. This polypeptide pattern was characterized by a single major protein band migrating with an apparent molecular weight of about 42,000 daltons which accounted for about 70% of the total protein on the gel. However, if the outer membrane protein is dissolved in SDS solution without urea and heated at 70 °C, major bands are observed in three regions of the gel: A broad band or group of bands near the top of the gel with an apparent molecular weight of much greater than 42,000 daltona (peak A), a second band with the same mobility as the 42,000-dalton band in boiled samples (peak B), and a third, faster-migrating band with an apparent molecular weight of less than 42,000 daltons (peak C).Elution of protein from A or C followed by heating at 100 °C converts this protein to a form migrating with peak B. If the outer-membrane protein is dissolved in SDS solution at 37 °C with no further heating and applied to gels, peak B dissappears completely and A and C increase. These can be partially converted to peak B by urea treatment. Protein from peaks A and C was isolated by chromatography on Sephadex in the presence of SDS, and the intrinsic viscosity of this protein was measured before and after boiling. The intrinsic viscosity of protein from peak A was 35 cc/g both before and after boiling, while the intrinsic viscosity of protein from peak C was 28 cc/g before boiling and 35 cc/g after boiling. These results are best explained by assuming that the protein in peak A represents aggregates of a 42,000-dalton species which are dissociated by boiling or solvent treatment and that the protein in peak C represents a monomeric form of the 42,000-dalton protein which is not fully reacted with SDS and which is converted to the “rigid rod” conformation characteristic of protein-SDS complexes only upon boiling or solvent treatment.  相似文献   

6.
Micelles of various surfactants containing protoporphyrin IX (PPIX) have been studied as models of primitive membrane-like photocatalytic systems. Spectral characteristics (absorption spectra, fluorescence emission and excitation spectra, fluorescence quantum yields and lifetimes) have been measured to indicate specific interactions of PPIX molecules with the micelles. Two types of PPIX aggregates are probably formed in water: non-fluorescent clusters corresponding to the absorption peak at 642-648 nm and fluorescent friable dimers with strong solute-solvent interactions corresponding to the absorption peak at 618 nm. The aggregates are solubilized by the micelles, which results in an increase of the fluorescence quantum yield (and thus in the increase of the PPIX sensitizing ability). Solubilization of PPIX molecules by SDS-micelles is enhanced by the partial neutralization of the negative surface charge of the micelles. Neutral Triton X-100 micelles solubilize PPIX much better than SDS particles; however, some of the clusters formed in the bulk aqueous phase of the detergent-water system remain aggregated in the crown of the micelle. The positively charged CTAB micelles have been shown to provide the best solubilization of PPIX accompanied by the highest increase of its emitting activity. The results are discussed in terms of the possible role of PPIX-containing membraneous systems in primitive photosynthesis.  相似文献   

7.
Ubiquitin has been purified to homogeneity, through a dialysis membrane having a NMW cutoff of 12 kDa, by taking advantage of its non-dialysable nature under these conditions. The dialysate was continuously recycled through a CM-52 cation exchange column at pH 4.5. The adsorbed fraction was eluted selectively at pH 7.2. Ubiquitin (25 mg) was obtained from 500 ml of packed RBCs. On SDS PAGE, ubiquitin showed varying mobility depending on the time of boiling in SDS. With 2 min of boiling, the molecular weight seemed to be 10.5 kDa, whereas 10 min of boiling resulted in a molecular weight of 8.5 kDa. Ubiquitin showed a slow intrinsic proteolytic activity against SDS-denatured beta-galactosidase in the absence of ATP. For the first 4 hr, there was no detectable degradation, but degradation was nearly complete after 8 hr. These data are not in agreement with those of Freid et al. [Proc. Natl Acad. Sci, USA, 84 (1987), 3685] who have reported a proteolytic activity comparable to that of other proteolytic enzymes.  相似文献   

8.
Cell wall glycoproteins of Candida albicans as released by different methods   总被引:10,自引:0,他引:10  
Different methods of extraction frequently used in other studies were used to release glycoproteins from both intact cells and isolated cell walls of yeast and hyphal forms of Candida albicans. Extracts were obtained from whole cells by treatment (i) with 2-mercaptoethanol (beta ME) at pH 8.6 and 37 C degrees and (ii) with zymolyase after treatment with beta ME. Extracts were obtained from isolated and washed cell walls (i) by boiling with beta ME and sodium dodecyl sulphate (SDS), (ii) by boiling with SDS and (iii) by treatment with zymolyase after SDS. The extracts were separated by SDS-polyacrylamide gel electrophoresis and analysed by Western blotting with four reagents. Analysis with concanavalin A (ConA) revealed different glycoprotein populations depending on the treatment. Three possible germ-tube-specific constituents were observed; and 80 kDa component released by beta ME from both intact cells and cell walls, and 47 kDa and 43 kDa moieties released by zymolyase only from intact cells. MAb 4C12, specific for the protein portion of a large germ tube constituent, recognized polydisperse material which just entered the gel in beta ME extracts and in the region extending up from 200 kDa to near the top of the gel in zymolyase extracts. MAb 24.17, specific for a carbohydrate determinant of yeast phase cells, reacted with disperse material in the region from the top of the gel to one-third to two-thirds the distance to the 220 kDa mass marker. Antiserum specific for the serotype A determinant of mannan reacted with large disperse component(s) migrating in the region from the top of the gel to about two-thirds the distance to the 220 kDa mass marker and with a 180 kDa component. The components recognized by MAb 4C12, but not those recognised by MAb 24.17 and serotype A antiserum, were effected by treatment with endo-beta-N-acetylglucosamidase H. The various analyses revealed that the method of extraction affected the composition and size of the constituents recognized by the reagents.  相似文献   

9.
The 12E7 murine monoclonal antibody recognizes a protease-sensitive component of human red cells, platelets and lymphocytes which could not be detected on granulocytes. Scatchard analyses indicated that the 125I-labelled antibody binds to 1000, 4000 and 27,000 antigen sites on each red cell, platelet and lymphocyte respectively, with a binding constant ranging from 4 x 10(7) to 9 x 10(7) M-1. The membrane components recognized by the monoclonal antibody were characterized by immunostaining on nitrocellulose sheets. A 28 kDa sialoglycoprotein was visualized following electrophoretic transfer of the red cell and lymphocyte membrane proteins separated by SDS/polyacrylamide-gel electrophoresis. Another component of 25 kDa was also clearly identified in the lymphocyte and platelet lysates, but was barely detectable in the red cell membrane preparations. Enzyme treatment of intact platelets, as well as analysis of the membrane and cytosolic preparations from these cells, have shown that the 25 kDa component was of cytoplasmic origin. The mobility of the 28 kDa membrane component is decreased following neuraminidase treatment of intact blood cells, but these cells still react normally with the monoclonal antibody, indicating that sialic acids are not required for binding. The 28 kDa component is present on red cell membranes prepared from S-s-U-, En(a-) and Gerbich(-) individuals, demonstrating that it is a new sialoglycoprotein not derived from glycophorins A, B, C or D. The 28 kDa component was totally solubilized with 0.1% Triton X-100 from red cell membranes and behaves like the other red cell membrane sialoglycoproteins since it was extracted in the aqueous phase following chloroform/methanol/water or butanol/water partitionings. The 28 kDa component could be partially purified by h.p.l.c. gel permeation chromatography and preparative SDS/polyacrylamide-gel electrophoresis. The material finally obtained strongly inhibits the 12E7 monoclonal as well as human anti-Xga antibodies, suggesting either that the 28 kDa glycoprotein carries both antigens or that the 12E7 and Xga-active molecules copurified.  相似文献   

10.
Previous studies have shown that human small-intestinal mucin consists of high-Mr glycoproteins and a smaller S-S-bonded protein of 118 kDa. The major antigenic determinants of the mucin were associated with the large glycoproteins, but depended for stability on intact disulphide bonds, and were destroyed by digestion with Pronase. In the present study we isolated and analysed the component parts of mucin from patients with cystic fibrosis with special attention being paid to the peptide constituents. After reduction with 0.2 M-beta-mercaptoethanol [5 min, 100 degrees C in 1% SDS (sodium dodecyl sulphate)], the large glycoproteins and smaller peptide with an apparent molecular size of 118 kDa were separated by equilibrium density-gradient centrifugation in CsCl, Sepharose 4B chromatography or preparative SDS/polyacrylamide-gel electrophoresis. The large glycoproteins contained about 70% of the protein of the native mucin. Digestion with Pronase resulted in a further loss of 'naked' protein (10% of the native mucin protein) from the C-terminal end of the glycoprotein peptide core, and left behind highly glycosylated proteins comprised mainly (70 mol%) of threonine, serine and proline. The 118 kDa component, which contained about 30% of the native mucin protein, consisted mainly of aspartic acid, serine, glutamic acid and glycine (40 mol%), plus threonine, proline, alanine, valine and leucine (35 mol%). Together with the 'naked' protein segment, the 118 kDa component contained most of the cysteine residues of the native mucin. Surprisingly, the peptide also contained carbohydrate (less than or equal to 5% of the native mucin carbohydrate but 50% by weight of the 118 kDa component), which included 9 mol% mannose, suggesting the presence of N-linked oligosaccharides. The peptide exhibited strong non-covalent interactions with the high-Mr glycoproteins and a tendency to self-aggregate in the absence of dissociating agents. Our findings therefore suggest that native mucin consists of large glycoproteins capable of forming disulphide bridges from their C-terminal 'naked' (antigenic) regions to a smaller glycopeptide having an Mr of 118 000.  相似文献   

11.
1. Rat liver glycosylasparaginase [N4-(beta-N-acetylglucosaminyl)-L-asparaginase, EC 3.5.1.26] was purified to homogeneity by using salt fractionation, CM-cellulose and DEAE-cellulose chromatography, gel filtration on Ultrogel AcA-54, concanavalin A-Sepharose affinity chromatography, heat treatment at 70 degrees C and preparative SDS/polyacrylamide-gel electrophoresis. The purified enzyme had a specific activity of 3.8 mumol of N-acetylglucosamine/min per mg with N4-(beta-N-acetylglucosaminyl)-L-asparagine as substrate. 2. The native enzyme had a molecular mass of 49 kDa and was composed of two non-identical subunits joined by strong non-covalent forces and having molecular masses of 24 and 20 kDa as determined by SDS/polyacrylamide-gel electrophoresis. 3. The 20 kDa subunit contained one high-mannose-type oligosaccharide chain, and the 24 kDa subunit had one high-mannose-type and one complex-type oligosaccharide chain. 4. N-Terminal sequence analysis of each subunit revealed a frayed N-terminus of the 24 kDa subunit and an apparent N-glycosylation of Asn-15 in the same subunit. 5. The enzyme exhibited a broad pH maximum above 7. Two major isoelectric forms were found at pH 6.4 and 6.6. 6. Glycosylasparaginase was stable at 75 degrees C and in 5% (w/v) SDS at pH 7.0.  相似文献   

12.
This paper describes the light microscopy (LM) and electron microscopy (EM) localization of synaptonemal complex (SC) antigens in oocytes of rats. For this purpose, we used monoclonal antibodies (Mabs) that recognize components of 30 + 33, 125, and 190 kDa antigens of SCs of rat spermatocytes. The LM localization was performed by immunofluorescence and the EM localization by immunogold staining. The reaction of the Mabs with oocytes was similar to the reaction with spermatocytes, but weaker. The 30 + 33 kDa as well as the 190 kDa antigens could always be demonstrated if axial elements of the SC were present, irrespective of whether these were paired or unpaired. Thus, these antigens could be detected from leptotene--early zygotene until diplotene. The 190-kDa antigen appeared in a diffuse manner just before the appearance of the 30 + 33 kDa antigens. The 30 + 33 kDa antigens were not only detected in the axial elements of SCs but also in characteristic aggregates, which appeared in zygotene and persisted until after the SCs had disappeared. Such aggregates had rarely been observed in spermatocytes. The 125 kDa antigen was only present in the tripartite segments of SCs, at the inner edge of the lateral elements. Thus, the reaction of the Mab against the 125 kDa antigen was detectable in zygotene, pachytene, and very early diplotene. It appeared later than 30 + 33 kDa and 190 kDa antigens and it disappeared earlier. We found that several steps of the immunostaining procedure could cause variation in the intensity of the Mab reaction.  相似文献   

13.
Lipopolysaccharide (LPS) bound to isolated porin was detected on polyacrylamide gels by using a carbohydrate-specific silver stain and on Western blots by using anti-lipid A monoclonal antibodies. Porin was isolated from Escherichia coli JF733 (Ra chemotype) and D21f2 (Re chemotype). Isolated porin was separated from loosely associated LPS by polyacrylamide gel electrophoresis (PAGE) in sodium dodecyl sulfate (SDS). Unheated porin traveled on gels as aggregates, presumably trimers, with an apparent molecular weight of 78,000 to 83,000. After heating to 100 degrees C for 2 min in SDS, the porin traveled as a monomer with a molecular weight of 36,000. The unheated, high-molecular-weight trimer band reacted in the gel with the carbohydrate-specific silver stain, while the heated monomer band showed no staining. In contrast, lipid A-specific monoclonal antibodies showed reactivity on Western blots to the 36,000-molecular-weight band but not to the trimer. Finally, both monomer and trimer bands were isolated from gels and rerun by SDS-PAGE. LPS was released from the trimer preparation when the sample was heated, but the monomer band that was formed by heating the trimer isolate still reacted with anti-lipid A antibodies. Quantitative Limulus amebocyte lysate analysis revealed an approximately equal molar ratio of LPS to protein in the electroeluted porin monomer. Thus, some but not all of the LPS could be released from trimer complexes by boiling in SDS. The isolated monomer did not release more LPS on boiling in SDS a second time but still had LPS tightly bound, as detected by lipid A-specific monoclonal antibodies.  相似文献   

14.
The primary leaves of young barley seedlings contain two major, extracellular, acid-soluble proteins of ca. 22 and 23 kDa apparent molecular mass. These proteins disappeared from the intercellular washing fluid upon stress treatments that enhanced H2O2 levels and that induced resistance to subsequent challenge by the powdery mildew fungus Erysiphe graminis f. sp. hordei. A partial peptide sequence of the 22 kDa protein was determined, and a cDNA clone was isolated. The 22 kDa protein belongs the the group of germin-like proteins (GLPs) and was designated HvGLP1. Despite its similarity to germin, i.e. oxalate oxidase, no oxalate oxidase activity of HvGLP1 could be detected. The RNA and soluble protein of HvGLP1 was highly abundant in young leaves, less abundant in older leaves and absent in roots. HvGLP1 RNA oscillated with a circadian rhythm, the minimum and maximum of RNA abundance being at the end of the dark and light periods, respectively. Heat and H2O2 treatment as well as pathogen infection caused disappearance of HvGLP1 protein from the fraction of soluble proteins of the intercellular space. HvGLP1 protein could be re-solubilized from cell walls of heat- or H2O2-treated leaves by boiling in SDS suggesting non-covalent cross linking. Although a physiological role of HvGLP1 insolubilization is still open, the protein may serve as marker for oxidative stress in cereals.  相似文献   

15.
We have previously noted that the Physicians' Desk Reference (PDR) contains over 80 instances in which a drug elicited a positive genotoxic response in one or more in vitro assays, despite having no obvious structural features predictive of covalent drug/DNA interactive potential or known mechanistic basis. Furthermore, in most cases, these drugs were "missed" by computational genotoxicity-predicting models such as DEREK, MCASE and TOPKAT. We have previously reported the application of a V79 cell-based model and a 3D DNA docking model for predicting non-covalent chemical/DNA interactions. Those studies suggested that molecules that are very widely structurally diverse may be capable of intercalating into DNA. To determine whether such non-covalent drug/DNA interactions might be involved in unexpected drug genotoxicity, we evaluated, using both models where possible, 56 marketed pharmaceuticals, 40 of which were reported as being clastogenic in in vitro cytogenetics assays (chromosome aberrations/mouse lymphoma assay). As seen before, the two approaches showed good concordance (62%) and 26 of the 40 (65%) drugs exhibiting in vitro clastogenicity were predicted as intercalators by one or both methods. This finding provides support for the hypothesis that non-covalent DNA interaction may be a common mechanism of clastogenicity for many drugs having no obvious structural alerts for covalent DNA interaction.  相似文献   

16.
With the lithium diiodosalicylate (LIS1) extraction-phenol partition method, we have isolated a sialoglycoprotein fraction from DBA/2 mouse erythrocyte ghosts. We have demonstrated that the Laemmeli system for SDS PAGE can resolve this fraction into four monomers of which two (gp- 2.1 and gp-3.1) appear to be authentic, whereas the other two (gp-2.2 and gp-3.2) are probably generated from gp-2.1 and gp-3.1, by limited proteolysis during the isolation procedure. All four components contain O-acetylated neuraminic acid residues, can be stained with Periodic acid-Schiff reagent (PAS) and with Coomassie Brilliant Blue (CB), and can be radioiodinated with the lactoperoxidase-glucose oxidase (LPO-GO) method. All monomers but especially gp-2.1 and gp-3.1 generate characteristic aggregates during solubilization in SDS. The aggregation is enhanced by boiling at high concentrations, and can be reversed by boiling at low concentrations. In addition, the fraction contains a diffuse component present also in ghosts which stains poorly with CB and with PAS and cannot be radioiodinated by the LPO-GO technique. SDS PAGE in the Steck and Yu gel system does not give an accurate separation of the sialoglycoprotein monomers.  相似文献   

17.
Following irradiation of bovine pancreatic ribonuclease in aqueous solution with 60Co gamma-rays protein aggregates are formed. The nature of the bonds linking these radiation-induced aggregates together has been investigated by chromatographic and electrophoretic methods. Thin-layer gel filtration and polyacrylamide gel electrophoresis, both in the presence of sodium dodecyl sulphate, demonstrated the existence of covalent crosslinks between the aggregates. However, non-covalent crosslinking also plays a role in the radiolysis of ribonuclease. Thin-layer gel filtration with and without 6 M urea and 2 per cent beta-mercaptoethanol added to the gel, revealed that only part of the covalent bonds between the aggregates consisted of disulphide linkages. By separation of the reduced aggregates by thin-layer gel filtration and electrophoresis, both with SDS, this finding was substantiated. Densitometric measurements indicated for example that the percentage of covalently linked dimers held together by disulphide bridges amounted to about 40-45 per cent, whereas the remaining 55-60 per cent of the dimers must be linked by other covalent bonds. The existence of covalent crosslinks other than disulphide bonds was also confirmed by isoelectric focusing. By this method definite differences were established between the proteolytic hydrolysates of the reduced aggregates and the reduced monomer of gamma-irradiated ribonuclease.  相似文献   

18.
A rapid in vitro assay was developed for monitoring protein-mediated cholesterol monomerization from bile acid aggregates. This assay uses a fluorescent cholesterol analog, 22-(N-(7-nitrobenz-2-oxa-1,3-diazol-4-yl)amino)-23,24-bisnor-5-cholen-3 beta-ol (NBD-cholesterol), which was shown to be absorbed by hamster in a fashion similar to cholesterol. The fluorescence of aggregates of NBD-cholesterol was strongly quenched in 2.5 mM of taurocholic acid. Addition of proteins from enterocytes of hamster small intestine led to a time- and dose-dependent dequenching of NBD-cholesterol fluorescence. Comparable dequenching can be detected with SDS and appears to involve monomerization of the NBD-cholesterol. Purification of enterocyte extract by sequential chromatography revealed an approximately 140-kDa protein complex (p140) able to mediate the monomerization of NBD-cholesterol. Each p140 complex mediated monomerization of 2.7 NBD-cholesterol molecules. The p140 complex appeared to be formed by dimerization of two approximately 58-kDa molecules since SDS-PAGE revealed a single dominant band at 58 kDa (p58). Protein sequence analysis suggested that p58 is protein-disulfide isomerase (PDI), and this conclusion was confirmed by cloning of hamster PDI, and detection of PDI enzyme activity in the purified fraction. Additional studies with either pure PDI or lysates of cells transfected with hamster PDI showed that PDI by itself was not sufficient for monomerizing cholesterol. Further, despite a similar mobility on SDS-PAGE (approximately 58 kDa), the p140 complex appeared approximately 45-kDa larger than pure PDI (approximately 95 kDa) when analyzed by a gel-filtration chromatography. The p140 complex may thus contain an unidentified molecule(s) in addition to PDI that may contribute importantly to cholesterol monomerization.  相似文献   

19.
Many bacterial pathogens present adhesins at the tips of long macromolecular filaments known as pili that are often important virulence determinants. Very little is known about how pili presented by Gram-positive pathogens mediate host cell binding. The crystal structure of a pilus adhesin from the important human pathogen Streptococcus pyogenes reveals an internal thioester bond formed between the side chains of a cysteine and a glutamine residue. The presence of the thioester was verified using UV-visible spectroscopy and mass spectrometry. This unusual bond has only previously been observed in thioester domains of complement and complement-like proteins where it is used to form covalent attachment to target molecules. The structure also reveals two intramolecular isopeptide bonds, one of these formed through a Lys/Asp residue pair, which are strategically positioned to confer protein stability. Removal of the internal thioester by allele-replacement mutagenesis in S. pyogenes severely compromises bacterial adhesion to model host cells. Although current paradigms of bacterial/host cell interaction envisage strong non-covalent interactions, the present study suggests cell adhesion could also involve covalent bonds.  相似文献   

20.
Very weak protein-protein interactions may play a critical role in cell physiology but they are not easily detectable in "in vitro" experiments. To detect these weak interactions, we have developed a strategy that included: (a) design of a rapid and very effective crosslinking of protein-protein complexes with poly-functional reagents; (b) selective adsorption of very large proteins on lowly activated ionic exchangers, based on the need of a multipoint physical adsorption to incorporate the proteins into the matrix; (c) purification by selective adsorption of protein-protein complexes formed by strong protein-protein interactions, via selective adsorption of the complexes on lowly activated ionic exchangers via multi-protein physical adsorption and leaving the non-associated proteins in the solution; (d) reinforcement of very weak protein-protein interactions by selective adsorption of the complex on lowly activated ionic exchange supports via a synergetic cooperation of the weak protein-protein interaction plus the interactions of both proteins with the support enabling the almost full shifting of the equilibrium towards the association position; (e) control of the aggregation state of proteins like BSA, formed by weak protein-protein interactions. In this last case, it seems that the interaction of the protein molecules placed on the borders of the aggregate with the groups on the support partially stabilizes the whole aggregate, although, some molecules of the aggregate cannot interact with the support. The size of the aggregates may be defined by controlling the concentration of ionised groups on the support: the less activated the supports are, the bigger the complexes. In this way, solid-phase proteomics could be a very interesting tool to detect weak protein-protein interactions.  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