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 0 毫秒
1.
Intracellular and secreted IgA from pokeweed mitogen (PWM)-stimulated normal peripheral blood lymphocytes, from 12-O-tetradecanoylphorbol-13-acetate (TPA)-stimulated peripheral blood lymphocytes of a patient with chronic lymphocytic leukemia (CLL), or from an IgA-producing human Epstein Barr virus (EBV)-transformed lymphoblastoid cell line were analyzed by molecular-sieve chromatography, electrophoresis in sodium dodecyl sulfate, and sucrose density ultracentrifugation. Fluorochrome-labeled anti-human IgA and secretory component (SC) were used as probes for the detection of polymeric IgA in individual cells. These methods demonstrated that the majority of intracellular IgA occurred in monomeric form, even when the predominant form of secreted IgA was polymeric. Sequential analyses of the IgA secreted by PWM-stimulated normal peripheral blood lymphocytes revealed that the proportion of polymeric IgA increased with the time of culture and that polymers represented the prevalent form of secreted IgA from the fifth day of culture. Although approximately one-half of TPA-stimulated CLL cells bound fluorochrome-labeled SC, only trace amounts of extracellular and intracellular polymeric IgA were detected in both culture supernatants and lysates. Culture supernatants of an IgA-secreting EBV-transformed cell line contained predominantly polymeric IgA. However, intracellular IgA was largely represented by monomers. The predominance of intracellular monomers in polymeric IgA-secreting cells suggested that the pathway of the assembly of human IgA molecules is analogous to that described for mouse IgA synthesis.  相似文献   

2.
Heme ligands were introduced in the hydrophobic core of an engineered monomeric ColE1 repressor of primer (rop-S55) in two different layers of the heptad repeat. Mutants rop-L63M/F121H (layer 1) and rop-L56H/L113H (layer 3) were found to bind heme with a K D of 1.1 ± 0.2 and 0.47 ± 0.07 μM, respectively. The unfolding of heme-bound and heme-free mutants, in the presence of guanidinium hydrochloride, was monitored by both circular dichroism and fluorescence spectroscopy. For the heme-bound rop mutants, the total free energy change was 0.5 kcal/mol higher in the layer 3 mutant compared with that in the layer1 mutant. Heme binding also stabilized these mutants by increasing the by 1.4 and 1.8 kcal/mol in rop-L63M/F121H and rop-L56H/L113H, respectively. The reduction potentials measured by spectroelectrochemical titrations were calculated to be −154 ± 2 mV for rop-56H/113H and −87.5 ± 1.2 mV for rop-L63M/F121H. The mutant designed to bind heme in a more buried environment (layer 3) showed tighter heme binding, a higher stability, and a different reduction potential compared with the mutant designed to bind heme in layer 1.  相似文献   

3.
Actobindin was previously shown to be an 88-residue polypeptide (Mr 9761) with an internal tandem repeat of 33-34 amino acids. Sedimentation equilibrium experiments have confirmed this Mr for native actobindin. Pyreneglyoxal-labeled actobindin had a similar Mr by sedimentation equilibrium analysis and bound to actin in a manner qualitatively similar to unmodified actobindin as determined by gel electrophoretic analysis of covalently cross-linked products. The stoichiometry of the actin-actobindin interaction was determined from the change in apparent Mr of pyrene-glyoxal-labeled actobindin in the presence of actin, as determined by scanning the ultracentrifuge cell at a wavelength that detected only the labeled protein. These data were consistent with the formation of a complex containing two actin and one actobindin molecules. The overall KD describing the binding of the first actin to either of the two sites on actobindin was 3.3 microM. The binding constant for the second actin suggested either negative cooperativity or inequality of the two actin-binding sites. Similar binding constants were obtained by analysis of the fluorescence enhancement that occurred when actobindin bound to actin labeled with either pyrene iodoacetamide or 4-(N-iodoacetoxyethyl-N-methyl)-7-nitrobenz-2-oxa-1,3-diazole. Cross-linking experiments with 1-ethyl-3-(3-dimethylaminopropyl)-carbodiimide and N-hydroxy-sulfosuccinimide qualitatively agreed with predictions made from a two-binding site model. Additionally, both the fluorescence and cross-linking experiments suggested that the interaction of the two actin molecules may contribute to the stability of the heterotrimeric complex.  相似文献   

4.
5.
Cooperativity is widespread in biology. It empowers a variety of regulatory mechanisms and impacts both the kinetic and thermodynamic properties of macromolecular systems. Traditionally, cooperativity is viewed as requiring the participation of multiple, spatially distinct binding sites that communicate via ligand-induced structural rearrangements; however, cooperativity requires neither multiple ligand binding events nor multimeric assemblies. An underappreciated manifestation of cooperativity has been observed in the non-Michaelis–Menten kinetic response of certain monomeric enzymes that possess only a single ligand-binding site. In this review, we present an overview of kinetic cooperativity in monomeric enzymes. We discuss the primary mechanisms postulated to give rise to monomeric cooperativity and highlight modern experimental methods that could offer new insights into the nature of this phenomenon. We conclude with an updated list of single subunit enzymes that are suspected of displaying cooperativity, and a discussion of the biological significance of this unique kinetic response.  相似文献   

6.
In order to clarify the catabolism sites of Hb-Hp and free Hb, the organ distributions of [125I]-Hb-Hp and [125I]-Hb were studied, and the cell types in each organ incorporating them were determined by immunohistochemical methods. After administration of [125I]-Hb-Hp in very small amounts to rats, 84.5% was incorporated into the liver, but the renal uptake was only 0.6%. [125I]-Hb was incorporated into the kidneys rather than into the liver when a fivefold greater amount of [125I]-Hb than the binding capacity of plasma Hp was administered. Parenchymal cells, but not Kupffer cells, in the liver were stained with anti-Hb or anti-Hp IgG after administration of Hb in an amount corresponding to the Hb binding capacity of Hp. The proximal tubule cells, but not the distal tubule cells, in the kidney were stained with anti-Hb IgG after administration of a fivefold greater amount of Hb than the binding capacity of Hp. On the basis of these results, we suggest that Hb-Hp was incorporated mainly into liver parenchymal cells and did not traverse glomeruli in the kidney. In contrast to Hb-Hp, free Hb could pass through the glomeruli easily and was incorporated into the proximal tubule cells.  相似文献   

7.
8.
Immunoglobulin A (IgA), the most abundant human immunoglobulin, mediates immune protection at mucosal surfaces as well as in plasma. It exists as two subclasses IgA1 and IgA2, and IgA2 is found in at least two allotypic forms, IgA2m(1) or IgA2m(2). Compared to IgA1, IgA2 has a much shorter hinge region, which joins the two Fab and one Fc fragments. In order to assess its solution structure, monomeric recombinant IgA2m(1) was studied by X-ray and neutron scattering. Its Guinier X-ray radius of gyration R(G) is 5.18 nm and its neutron R(G) is 5.03 nm, both of which are significantly smaller than those for monomeric IgA1 at 6.1-6.2 nm. The distance distribution function P(r)for IgA2m(1) showed a broad peak with a subpeak and gave a maximum dimension of 17 nm, in contrast to the P(r) curve for IgA1, which showed two distinct peaks and a maximum dimension of 21 nm. The sedimentation coefficients of IgA1 and IgA2m(1) were 6.2S and 6.4S, respectively. These data show that the solution structure of IgA2m(1) is significantly more compact than IgA1. The complete monomeric IgA2m(1) structure was modelled using molecular dynamics to generate random IgA2 hinge structures, to which homology models for the Fab and Fc fragments were connected to generate 10,000 full models. A total of 104 compact best-fit IgA2m(1) models gave good curve fits. These best-fit models were modified by linking the two Fab light chains with a disulphide bridge that is found in IgA2m(1), and subjecting these to energy refinement to optimise this linkage. The averaged solution structure of the arrangement of the Fab and Fc fragments in IgA2m(1) was found to be predominantly T-shaped and flexible, but also included Y-shaped structures. The IgA2 models show full steric access to the two FcalphaRI-binding sites at the Calpha2-Calpha3 interdomain region in the Fc fragment. Since previous scattering modelling had shown that IgA1 also possessed a flexible T-shaped solution structure, such a T-shape may be common to both IgA1 and IgA2. The final models suggest that the combination of the more compact IgA2m(1) and the more extended IgA1 structures will enable human IgA to access a broader range of antigens than either acting alone. The hinges of both IgA subclasses appear to show reduced flexibility when compared to their equivalents in IgG, and this may be important for maintaining an extended IgA structure.  相似文献   

9.
Rat serum albumin was labeled with [G-3H]raffinose, a nondegradable radioactive tracer which has been shown to accumulate in cells following protein degradation (Van Zile et al., 1979, J. Biol. Chem.254, 3547–3553). Stoichiometric labeling of albumin with [3H]raffinose did not affect either its circulating half-life or its equilibration into extravascular spaces. The tissue distribution of acid-soluble radioactivity was followed for up to 96 h after intracardial injection of labeled protein. At 18, 48, and 96 h postinjection muscle and hide accounted for the highest proportion (40–60%) of catabolized dose of albumin in the body. Additional small percentages of radioactive degradation products were recovered in liver, kidney, and other visceral organs. The data provide the first direct evidence that the major fraction of albumin catabolism in the rat occurs in muscle and hide.  相似文献   

10.
In J774 murine macrophages, chemically oxidized LDL (OxLDL) and biologically oxidized LDL (BioOxLDL) have similar metabolic fates, characterized by a relatively poor degradation when compared with acetylated LDL (AcLDL), and a modest ability to activate acyl-CoA:cholesterol acyltransferase (ACAT) (850 and 754 pmol [14C]oleate/mg cell protein in OxLDL- and BioOxLDL-incubated cells, versus 425 and 7070 pmol [14C]cholesteryl oleate/mg cell protein in control and AcLDL-incubated cells) with a massive increase of cellular free cholesterol. Therefore, OxLDL were used to investigate the cellular processing of oxidatively modified LDL. Binding and fluorescence microscopy studies demonstrated that OxLDL are effectively bound and internalized by macrophages and accumulate in organelles with density properties similar to those of endo/lysosomes. Although the overall metabolism of OxLDL is modestly affected by 100 microM chloroquine, owing to the poor cellular degradation of the substrate, the drug can further depress OxLDL degradation, indicating that this process takes place in an acidic compartment. Failure to detect products of extensive degradation of OxLDL in the medium is due to their relative resistance to enzymatic hydrolysis, as demonstrated also by in vitro experiments with partially purified lysosomal enzymes, rather than to the intracellular accumulation of degradation products (degraded intracellular protein is, at most, 8.5% of total). This sluggish degradation process is not due to a cytotoxic effect since OxLDL do not affect the intracellular processing of other ligands like AcLDL or IgG. The accumulation of OxLDL-derived products within macrophages may elicit cellular responses, the relevance of which in the atherosclerotic process remains to be addressed.  相似文献   

11.
Depositions of IgA in the renal glomerular mesangial area are a hallmark of IgA nephropathy, and are thought to be crucial for the onset of inflammation processes in IgA nephropathy. In this report we show that human mesangial cells (MC) in vitro bind IgA and that binding of IgA enhances the production of IL-6 by MC. Furthermore we show that the size of IgA is crucial in its capability to enhance IL-6 production. Monomeric IgA does not affect basic IL-6 production, whereas dimeric and polymeric IgA enhance IL-6 production up to 3- to 9-fold respectively. Additional studies demonstrate that enhanced IL-6 production by MC is not accompanied by increased proliferation of human mesangial cells, a finding which is distinct from that found with rat mesangial cells. Taken together, these fmdings suggest that deposition of dimeric and polymeric IgA in the mesangial area of human kidneys in IgA nephropathy may amplify local inflammation.  相似文献   

12.
Yeast glyoxalase I is a monomeric enzyme with two active sites   总被引:3,自引:0,他引:3  
The tertiary structure of the monomeric yeast glyoxalase I has been modeled based on the crystal structure of the dimeric human glyoxalase I and a sequence alignment of the two enzymes. The model suggests that yeast glyoxalase I has two active sites contained in a single polypeptide. To investigate this, a recombinant expression clone of yeast glyoxalase I was constructed for overproduction of the enzyme in Escherichia coli. Each putative active site was inactivated by site-directed mutagenesis. According to the alignment, glutamate 163 and glutamate 318 in yeast glyoxalase I correspond to glutamate 172 in human glyoxalase I, a Zn(II) ligand and proposed general base in the catalytic mechanism. The residues were each replaced by glutamine and a double mutant containing both mutations was also constructed. Steady-state kinetics and metal analyses of the recombinant enzymes corroborate that yeast glyoxalase I has two functional active sites. The activities of the catalytic sites seem to be somewhat different. The metal ions bound in the active sites are probably one Fe(II) and one Zn(II), but Mn(II) may replace Zn(II). Yeast glyoxalase I appears to be one of the few enzymes that are present as a single polypeptide with two active sites that catalyze the same reaction.  相似文献   

13.
Specific cDNA probes of Moloney and AKR murine leukemia viruses have been prepared to characterize the proviral integration sites of these viruses in the genomes of Balb/Mo and Balb/c mice. The genetically transmitted Moloney provirus of Balb/Mo mice was detected in a characteristic Eco RI DNA fragment of 16 x 10(6) daltons. No fragment of this size was detected in tissue DNAs from Balb/c mice infected as newborns with Moloney virus. We conclude that a viral integration site, occupied in preimplantation mouse embryos, is not necessarily occupied when virus infects cells in post-natal animals. Balb/Mo and Balb/c mice do carry the AkR structural gene in an Eco RI DNA fragment of 12 x 10(6) daltons. Further restriction analysis of this fragment indicated that both mouse lines carry one AKR-type provirus. Leukemogenesis in Balb/Mo and newborn infected Balb/c mice is accompanied by reintegration of Moloney viral sequences in new chromosomal sites of tumor tissues. Part of the reintegrated Moloney viral sequences are of subgenomic size. The AKR viral sequences, however, are not found in new sites. Further restriction analysis revealed that the development of Moloney virus-induced leukemia in Balb/Mo mice does not lead to detectable structural alteration of the genetically transmitted Moloney and AKR structural genes. Possible mechanisms of the reintegration process are also discussed.  相似文献   

14.
Binding characteristics of the galactose/N-acetylgalactosamine-specific, hepatic lectins of rabbit and rat were studied using small, high-affinity ligands containing two and three N-acetylgalactosamine residues per molecule [Lee, R. T. and Lee, Y. C. (1987) Glycoconjugate J. 4, 317-328]. These N-acetylgalactosamine cluster ligands have the receptor-ligand dissociation constants in nanomolar range, so that the lectin-ligand interaction can easily studied by an equilibrium (gel chromatography) or non-equilibrium (fast filtration assay) method. The results suggest that there exist on the average two N-acetylgalactosamine-combining sites per monomeric unit of both the rabbit and rat lectins.  相似文献   

15.
Acetylation of Ser-530 of sheep prostaglandin endoperoxide (PGG/H) synthase by aspirin causes irreversible inactivation of the cyclooxygenase activity of the enzyme. To determine the catalytic function of the hydroxyl group of Ser-530, we used site-directed mutagenesis to replace Ser-530 with an alanine. Cos-1 cells transfected with expression vectors containing the native (Ser-530) or mutant (Ala-530) cDNAs for sheep PGG/H synthase expressed comparable cyclooxygenase and hydroperoxidase activities. Km values for arachidonate (8 microM) and ID50 values for reversible inhibition by the cyclooxygenase inhibitors, flurbiprofen (5 microM), flufenamate (20 microM), and aspirin (20 mM), were also the same for both native and mutant PGG/H synthases; however, only the native enzyme was irreversibly inactivated by aspirin. Thus, the "active site" Ser-530 of PGG/H synthase is not essential for catalysis or substrate binding. Apparently, acetylation of native PGG/H synthase by aspirin introduces a bulky sidechain at position 530 which interferes with arachidonate binding. In related studies, a cDNA for mouse PGG/H synthase was cloned and sequenced. A sequence of 35 residues with Ser-530 at the midpoint was identical in the two proteins. Thus, Ser-530 does lie in a highly conserved region, probably involved in cyclooxygenase catalysis. Sequence comparisons of mouse and sheep PGG/H synthase also provided information about the heme-binding site of the enzyme. The sheep HYPR sequence (residues 274-277), which had been proposed to form a portion of the distal heme-binding site, is not conserved in the mouse PGG/H synthase, suggesting that this region is not the distal heme-binding site. One sequence, TIWLREHNRV (residues 303-312 of the sheep enzyme), is very closely related to the sequence TLW(L)LREHNRL common to thyroid peroxidase and myeloperoxidase. The histidine in this latter sequence is the putative axial heme ligand of these peroxidases. We suggest that the histidine (His-309) of sheep PGG/H synthase sequence is the axial heme ligand of this enzyme.  相似文献   

16.
M cells represent the primary route by which mucosal Ags are transported across the intestinal epithelium and delivered to underlying gut-associated lymphoid tissues. In rodents and rabbits, Peyer's patch M cells selectively bind and endocytose secretory IgA (SIgA) Abs. Neither the nature of the M cell IgR nor the domains of SIgA involved in this interaction are known. Using a mouse ligated ileal loop assay, we found that monoclonal IgA Abs with or without secretory component, but not IgG or IgM Abs, bound to the apical surfaces of Peyer's patch M cells, indicating that the receptor is specific for the IgA isotype. Human serum IgA and colostral SIgA also bound to mouse M cells. The asialoglycoprotein receptor or other lectin-like receptors were not detected on the apical surfaces of M cells. We used recombinant human IgA1 and human IgA2 Abs and domain swapped IgA/IgG chimeras to determine that both domains Calpha1 and Calpha2 are required for IgA adherence to mouse Peyer's patch M cells. This distinguishes the M cell IgA receptor from CD89 (FcalphaI), which binds domains Calpha2-Calpha3. Finally, we observed by immunofluorescence microscopy that some M cells in the human ileum are coated with IgA. Together these data suggest that mouse, and possibly human, M cells express an IgA-specific receptor on their apical surfaces that mediates the transepithelial transport of SIgA from the intestinal lumen to underlying gut-associated organized lymphoid tissues.  相似文献   

17.
IgA-induced avidity maturation of IgA Fc receptors on murine T lymphocytes   总被引:2,自引:0,他引:2  
The analysis of 30 well characterized murine T lymphocyte populations using a cytofluorometric IgA binding assay has identified many populations that are constitutive and/or inducible for IgA receptor expression, and has identified two distinct mechanisms by which IgA up-regulates the IgA-binding properties of murine T cells. Studies with lymphomas, hybridomas, Ag-specific clones and activated normal splenic T cells identified many examples of CD4 and CD8 lineage cells that constitutively expressed IgA receptors. T cell populations that constitutively expressed IgA receptors exhibited enhanced IgA binding after incubation with oligomeric IgA for 18 h. The IgA-induced up-regulation of IgA binding resulted from two distinct processes: 1) an increase in the number of surface membrane IgA binding sites and 2) an increase in the avidity of IgA binding without a change in the number of binding sites. The IgA-induced avidity increase was reflected by a 5- to 10-fold decrease in the apparent Kd. Depending on the T cell population examined the enhanced binding of IgA involved one or both of these mechanisms. T cell populations that did not constitutively express IgA receptors failed to bind IgA after prolonged incubation with oligomeric IgA suggesting that if such cells can express IgA receptors they require other signals to induce their expression. Consistent with this possibility is the finding that resting splenic T cells did not bind IgA but their activation with Con A or mAb anti-T3 resulted in high level expression of IgA receptors. These studies have identified multiple distinct mechanisms that alter the IgA-binding properties of murine T cells and are discussed in terms of their possible physiologic significance.  相似文献   

18.
G J Roth  K Titani  L W Hoyer  M J Hickey 《Biochemistry》1986,25(26):8357-8361
Purified human plasma von Willebrand factor (vWf) binds to pepsin-digested monomeric type III collagen in a saturable (KD = 1 X 10(-8) M), specific, and rapid manner with a stoichiometry of approximately 1:15 [vWf subunit (Mr 270,000):collagen trimer (Mr 300,000)]. Two reduced and alkylated CNBr peptides of vWf, termed M11 residues 542-622 and M20 residues 948-998 [Titani, K., Kumar, S., Takio, K., Ericsson, L. H., Wade, R. D., Ashida, K., Walsh, K. A., Chopek, M. W., Sadler, J. E., & Fujikawa, K. (1986) Biochemistry 25, 3171-3184], inhibited vWf binding to collagen. With 125I-vWf (2 X 10(-9) M) as ligand, M11, M20, fragment III (a dimeric, V8 protease, NH2-terminal fragment, Mr 320,000 referenced above), and unlabeled vWf inhibited binding to collagen with EC50 values of 4.8 X 10(-7), 9.4 X 10(-7), 1.1 X 10(-7), and 0.2 X 10(-7) M, respectively. M11 and M20 bind to collagen directly when 125I-labeled peptides are used as ligands. Other CNBr fragments of vWf were less effective as inhibitors (5-fold or less) and bound less avidly to collagen (5-fold or less) compared to M11 and M20. A murine anti-human vWf monoclonal antibody (MR5), which blocks the binding of vWf to collagen, bound selectively to both M11 and M20 when tested in an enzyme-linked immunoadsorbent assay.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Glyoxalase I (GloI) catalyzes the glutathione-dependent conversion of 2-oxoaldehydes to S-2-hydroxyacylglutathione derivatives. Studies on GloI from diverse organisms such as man, bacteria, yeast, and different parasites show striking differences among these potentially isofunctional enzymes as far as metal content and the number of active sites per subunit are concerned. So far, it is not known whether this structural variability is linked to catalytic or regulatory features in vivo. Here we show that recombinant GloI from the malaria parasite Plasmodium falciparum has a high- and a low-affinity binding site for the diastereomeric hemithioacetals formed by addition of glutathione to methylglyoxal. Both active sites of the monomeric enzyme are functional and have similar k(cat)(app) values. Proteolytic susceptibility studies and detailed analyses of the steady-state kinetics of active-site mutants suggest that both reaction centers can adopt two discrete conformations and are allosterically coupled. As a result of the positive homotropic allosteric coupling, P. falciparum GloI has an increased affinity at low substrate concentrations and an increased activity at higher substrate concentrations. This could also be the case for GloI from yeast and other organisms. Potential physiologically relevant differences between monomeric GloI and homodimeric GloI are discussed. Our results provide a strong basis for drug development strategies and significantly enhance our understanding of GloI kinetics and structure-function relationships. Furthermore, they extend the current knowledge on allosteric regulation of monomeric proteins in general.  相似文献   

20.
Sequence of the VH gene for murine IgA X-24.   总被引:2,自引:1,他引:1       下载免费PDF全文
  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