首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Calcitonin receptor-like receptor (CRLR) constitutes either a CGRP receptor when complexed with receptor activity-modifying protein 1 (RAMP1) or an adrenomedullin receptor when complexed with RAMP2 or RAMP3. RAMP proteins modify the glycosylation status of CRLR and determine their receptor specificity; when treated with tunicamycin, a glycosylation inhibitor, CHO-K1 cells constitutively expressing both RAMP2 and CRLR lost the capacity to bind adrenomedullin. Similarly, in HEK293 EBNA cells constitutively expressing RAMP1/CRLR receptor complex CGRP binding was remarkably inhibited. Whichever RAMP protein was co-expressing with CRLR, the ligand binding was sensitive to tunicamycin. There are three putative Asn-linked glycosylation sites in the extracellular, amino terminal domain of CRLR at positions 66, 118 and 123. Analysis of CRLR mutants in which Gln was substituted for selected Asn residues showed that glycosylation of Asn123 is required for both the binding of adrenomedullin and the transduction of its signal. Substituting Asn66 or Asn118 had no effect. FACS analysis of cells expressing FLAG-tagged CRLRs showed that disrupting Asn-linked glycosylation severely affected the transport of the CRLR protein to the cell surface on N66/118/123Q mutant, and slightly reduced the level of the cell surface expression of N123Q mutant compared with wild-type CRLR. But other single mutants (N66Q, N118Q) had no effect for other single mutants. Our data shows that glycosylation of Asn66 and Asn118 is not essential for ligand binding, signal transduction and cell surface expression, and Asn123 is important for ligand binding and signal transduction rather than cell surface expression. It thus appears that glycosylation of Asn123 is required for CRLR to assume the appropriate conformation on the cell surface through its interaction with RAMPs.  相似文献   

2.
Plasma sex hormone-binding globulin (SHBG) and testicular androgen-binding protein (ABP) are homodimeric glycoproteins that share the same primary structure, and differ only with respect to the types of oligosaccharides associated with them. The biological significance of these differences is not understood, but enzymatically deglycosylated SHBG and a non-glycosylated SHBG mutant both bind steroids normally. Various affinity-labelling experiments, and studies of recombinant SHBG mutants have indicated that a region encompassing and including Met-139 in human SHBG represents an important component of its steroid-binding site. Analyses of chimeric proteins comprising various portions of human SHBG and rat ABP have also indicated that residues important for the much higher affinity of human SHBG for steroid ligands are probably located within the N-terminal portion of these molecules. Recent studies of SHBG mutants have confirmed this, and a deletion mutant containing only the first 205 N-terminal residues of human SHBG has been produced which dimerizes and binds steroids appropriately. The introduction of amino-acid substitutions between Lys-134 and Phe-148 of SHBG has also indicated that residues including and immediately N-terminal of Met-139 may influence steroid-binding specificity, while those immediately C-terminal of Met-139 represent at least a part of the dimerization domain. These studies have also demonstrated that dimerization is induced by the presence of steroid ligand in the binding site, and that divalent cations play an important role in this process. Together, these data have led us to conclude that SHBG is a modular protein, which comprises an N-terminal steroid-binding and dimerization domain, and a C-terminal domain containing a highly-conserved consensus sequence for glycosylation that may be required for other biological activities, such as cell-surface recognition.  相似文献   

3.
Tissue factor pathway inhibitor (TFPI) produced by endothelial cells contains sulfated Asn-linked oligosaccharides. We have determined that greater than 70% of the oligosaccharides on recombinant TFPI expressed in 293 cells terminate with the sequence SO4-4GalNAc beta 1, 4GlcNAc beta 1, 2Man alpha. Oligosaccharides terminating with this sequence have previously been described on lutropin, thyrotropin, and pro-opiomelanocortin: glycoproteins synthesized in the anterior pituitary. A GalNAc-transferase that recognizes the tripeptide motif Pro-Xaa-Arg/Lys 6-9 residues N-terminal to Asn glycosylation sites accounts for the specific addition of GalNAc to the oligosaccharide acceptor on these glycoproteins, whereas a GalNAc beta 1,4GlcNAc beta 1, 2Man alpha-4-sulfotransferase accounts for the addition of sulfate. The sulfated oligosaccharides present on these hormones are responsible for their rapid clearance from plasma by a receptor in hepatic reticuloendothelial cells. GalNAc- and sulfotransferase activities with the same properties as those expressed in the pituitary are detected at high levels in 293 cells and at lower levels in endothelial cells. Chinese hamster ovary (CHO) cells do not contain detectable levels of either transferase and rTFPI expressed in CHO cells does not contain sulfated Asn-linked oligosaccharides. TFPI contains the sequence Pro-Phe-Lys, 9 residues N-terminal to the glycosylation site at position 228; this tripeptide may act as the recognition sequence for the GalNAc-transferase. rTFPI produced by 293 cells, but not that produced by CHO cells, is bound by the receptor on hepatic reticuloendothelial cells suggesting the sulfated structures play a role in the biologic behavior of TFPI.  相似文献   

4.
5.
The hemagglutinin (HA) of fowl plague virus A/FPV/Rostock/34 (H7N1) carries two N-linked oligosaccharides attached to Asn123 and Asn149 in close vicinity to the receptor-binding pocket. In previous studies in which HA mutants lacking either one (mutants G1 and G2) or both (mutant G1,2) glycosylation sites had been expressed from a simian virus 40 vector, we showed that these glycans regulate receptor binding affinity (M. Ohuchi, R. Ohuchi, A. Feldmann, and H. D. Klenk, J. Virol. 71:8377-8384, 1997). We have now investigated the effect of these mutations on virus growth using recombinant viruses generated by an RNA polymerase I-based reverse genetics system. Two reassortants of influenza virus strain A/WSN/33 were used as helper viruses to obtain two series of HA mutant viruses differing only in the neuraminidase (NA). Studies using N1 NA viruses revealed that loss of the oligosaccharide from Asn149 (mutant G2) or loss of both oligosaccharides (mutant G1,2) has a pronounced effect on virus growth in MDCK cells. Growth of virus lacking both oligosaccharides from infected cells was retarded, and virus yields in the medium were decreased about 20-fold. Likewise, there was a reduction in plaque size that was distinct with G1,2 and less pronounced with G2. These effects could be attributed to a highly impaired release of mutant progeny viruses from host cells. In contrast, with recombinant viruses containing N2 NA, these restrictions were much less apparent. N1 recombinants showed lower neuraminidase activity than N2 recombinants, indicating that N2 NA is able to partly overrule the high-affinity binding of mutant HA to the receptor. These results demonstrate that N-glycans flanking the receptor-binding site of the HA molecule are potent regulators of influenza virus growth, with the glycan at Asn149 being dominant and that at Asn123 being less effective. In addition, we show here that HA and NA activities need to be highly balanced in order to allow productive influenza virus infection.  相似文献   

6.
Recombinant human tissue plasminogen activator (rt-PA), produced by expression in Chinese hamster ovary cells, is a fibrin-specific plasminogen activator which has been approved for clinical use in the treatment of myocardial infarction. In this study, the structures of the Asn-linked oligosaccharides of Chinese hamster ovary-expressed rt-PA have been elucidated. High mannose and hybrid oligosaccharides were released from the protein by endoglycosidase H digestion, whereas N-acetyllactosamine-type ("complex") oligosaccharides were released by peptide:N-glycosidase F digestion. The oligosaccharides were fractionated by gel permeation chromatography and anion exchange high performance liquid chromatography (HPLC), and their structures were analyzed by composition and methylation analysis, high pH anion exchange chromatography, fast atom bombardment-mass spectrometry (FAB-MS), and 500-MHz 1H NMR spectroscopy. High mannose oligosaccharides were found to account for 38% of the total carbohydrate content of rt-PA and consisted of Man5GlcNAc2, Man6GlcNAc2, and Man7GlcNAc2 in the ratio 1.8:1.7:1. Two hybrid oligosaccharides were identified and accounted for 3% of the carbohydrate of rt-PA. The N-acetyllactosamine-type oligosaccharides were found to comprise diantennary (34% of total carbohydrate), 2,4-branched triantennary (11%), 2,6-branched triantennary (9%), and tetraantennary (5%) structures. Sialylation of these oligosaccharides was by alpha (2----3) linkages to galactose. Most (greater than 90%) of the N-acetyllactosamine-type structures contained fucose alpha (1----6) linked to the Asn-linked N-acetylglucosamine residue. The distribution of oligosaccharide structures at individual glycosylation sites (Asn residues 117, 184, and 448) was also determined. rt-PA exists as two variants that differ by the presence (type I) or absence (type II) of carbohydrate at Asn-184. Tryptic glycopeptides were isolated by reversed phase high performance liquid chromatography and treated with peptide:N-glycosidase F. The oligosaccharides released from each glycosylation site were analyzed by high pH anion exchange chromatography. By this analysis, Asn-117 was demonstrated to carry exclusively high mannose oligosaccharides. When glycosylated, Asn-184 carried diantennary, 2,4-branched triantennary, 2,6-branched triantennary, and tetraantennary N- acetyllactosamine oligosaccharides in the ratio 9.0:4.5:1.4:1. Asn- 448 carried the same types of oligosaccharides, but in the ratio 7.5:1.6:2.1:1. The distributions of Asn-linked oligosaccharides at positions 117 and 448 were found not to be affected by the presence or absence of carbohydrate at position 184. The relevance of the  相似文献   

7.
We have reported that bovine DNase I, a secretory glycoprotein, acquires mannose 6-phosphate residues on 12.6% of its Asn-linked oligosaccharides when expressed in COS-1 cells and that the extent of phosphorylation increases to 79.2% when lysines are placed at positions 27 and 74 of the mature protein (Nishikawa, A., Gregory, W. , Frenz, J., Cacia, J., and Kornfeld, S. (1997) J. Biol. Chem. 272, 19408-19412). We now demonstrate that murine DNase I, which contains Lys27 and Lys74, is phosphorylated only 20.9% when expressed in the same COS-1 cell system. This difference is mostly due to the absence of three residues present in bovine DNase I (Tyr54, Lys124, and Ser190) along with the presence of a valine at position 23 that is absent in the bovine species. We show that Val23 inhibits phosphorylation at the Asn18 glycosylation site, whereas Tyr54, Lys124, and Ser190 enhance phosphorylation at the Asn106 glycosylation site. Tyr54 and Ser190 are widely separated from each other and from Asn106 on the surface of DNase I, indicating that residues present over a broad area influence the interaction with UDP-GlcNAc:lysosomal enzyme N-acetylglucosamine-1-phosphotransferase, which is responsible for the formation of mannose 6-phosphate residues on lysosomal enzymes.  相似文献   

8.
A role for N-linked oligosaccharides on the biochemical properties of recombinant alpha-l-arabinofuranosidase 54 (AkAbf54) defined in glycoside hydrolase family 54 from Aspergillus kawachii expressed in Pichia pastoris was analyzed by site-directed mutagenesis. Two N-linked glycosylation motifs (Asn(83)-Thr-Thr and Asn(202)-Ser-Thr) were found in the AkAbf54 sequence. AkAbf54 comprises two domains, a catalytic domain and an arabinose-binding domain classified as carbohydrate-binding module 42. Two N-linked glycosylation sites are located in the catalytic domain. Asn(83), Asn(202), and the two residues together were replaced with glutamine by site-directed mutagenesis. The biochemical properties and kinetic parameters of the wild-type and mutant enzymes expressed in P. pastoris were examined. The N83Q mutant enzyme had the same catalytic activity and thermostability as the wild-type enzyme. On the other hand, the N202Q and N83Q/N202Q mutant enzymes exhibited a considerable decrease in thermostability compared to the glycosylated wild-type enzyme. The N202Q and N83Q/N202Q mutant enzymes also had slightly less specific activity towards arabinan and debranched arabinan. However, no significant effect on the affinity of the mutant enzymes for the ligands arabinan, debranched arabinan, and wheat and rye arabinoxylans was detected by affinity gel electrophoresis. These observations suggest that the glycosylation at Asn(202) may contribute to thermostability and catalysis.  相似文献   

9.
The human calcitonin (CT) receptor-like receptor (hCRLR) of the B family of G protein-coupled receptors is N-glycosylated and associates with receptor-activity-modifying proteins for functional interaction with CT gene-related peptide (CGRP) or adrenomedullin (ADM), respectively. Three putative N-glycosylation sites Asn(60), Asn(112) and Asn(117) are present in the amino-terminal extracellular domain of the hCRLR. Tunicamycin dose-dependently inhibited the glycosylation of a myc-tagged hCRLR and in parallel specific [(125)I]CGRP and -ADM binding. Similarly, the double mutant myc-hCRLR(N60,112T) exhibited minimal N-glycosidase F sensitive glycosylation, presumably at the third Asn(117), and the cell surface expression and specific radioligand binding were impaired. Substitution of the Asn(117) by Thr abolished CGRP and ADM binding in the face of intact N-glycosylation and cell surface expression.  相似文献   

10.
Human chorionic gonadotropin (hCG), luteinizing hormone (LH), follicle-stimulating hormone and thyroid-stimulating hormone are a family of glycoprotein hormones that share a common alpha subunit but differ in their hormone-specific beta subunits. Using site-directed mutagenesis and gene-transfer, we analyzed the role of the N-linked oligosaccharides of alpha and chorionic gonadotropin (CG)beta in the secretion, assembly, and biologic activity of hCG. Absence of carbohydrate at alpha asparagine (Asn) 52 decreased combination with CG beta but did not alter monomer secretion. Absence of the alpha Asn78 oligosaccharide increased the degradation of the alpha subunit, but the presence of CG beta stabilized this alpha mutant in an efficiently formed dimer complex. Alternatively, absence of both alpha oligosaccharides slowed both secretion and dimer formation but allowed an intermediate level of alpha secreted or dimerized compared to the single-site mutants. Analysis of the CG beta glycosylation mutants revealed that absence of the Asn30 oligosaccharide, but not Asn13, slowed secretion but not assembly, whereas absence of both oligosaccharides slowed both secretion and dimer formation. Analysis of the receptor binding of the hCG glycosylation mutants showed that absence of any or all of the hCG N-linked oligosaccharides had only a minor effect on receptor affinity of the derivatives. However, the absence of alpha Asn52, but not the alpha Asn78 or the CG beta carbohydrate units, reduced the steroidogenic effect, unmasked differences in the beta oligosaccharides, and converted the deglycosylated derivatives into antagonists.  相似文献   

11.
The structures of the oligosaccharides of the hemagglutinin of fowl plague virus [influenza A/FPV/Rostock/34 (H7N1)] have been elucidated by one- and two-dimensional 1H n.m.r. spectroscopy at 500 MHz and by microscale methylation analysis. N-Glycosidic oligosaccharides of the oligomannosidic (OM) and of the N-acetyllactosaminic type have been found, the latter type comprising biantennary structures, without (A) or with (E) bisecting N-acetylglucosamine, and triantennary (C) structures. Analysis of the tryptic and thermolytic glycopeptides of the hemagglutinin allowed the allocation of these oligosaccharides to the individual glycosylation sites. Each attachment site contained a unique set of oligosaccharides. Asn12 contains predominantly structures C and E which are highly fucosylated. Asn28 contains OM and A structures that lack fucose and sulfate. Asn123 shows A that has incomplete antennae but is highly fucosylated and sulfated. Asn149 has fucosylated A and E. Asn231 shows fucosylated A and E with incomplete antennae. Asn406 has OM oligosaccharides. Asn478 has A and E with little fucose. Localization of the oligosaccharides on the three-dimensional structure of the hemagglutinin revealed that the oligomannosidic glycans are attached to glycosylation sites at which the enzymes responsible for carbohydrate processing do not have proper access. These observations demonstrate that an important structural determinant for the oligosaccharide side chains is the structure of the glycoprotein itself. In addition, evidence was obtained that the rate of glycoprotein synthesis also has an influence on carbohydrate structure.  相似文献   

12.
Human serum macrophage-stimulating protein (MSP) induces motile activity of murine resident peritoneal macrophages and is a growth and motility factor for epithelial cells. It belongs to the plasminogen-related family of kringle proteins, and is secreted as a single-chain, 78-kDa, biologically inactive pro-MSP. Proteolytic cleavage of pro-MSP at a single site yields active MSP, a disulfide-linked alphabeta-chain heterodimer. However cleavage of recombinant pro-MSP yielded not only the disulfide-linked heterodimer, but also free alpha- and beta-chains, indicating that some of the recombinant molecules lacked an alphabeta-chain disulfide. We purified the free chains for characterization. The beta-chain of MSP has three extra cysteines, Cys527, Cys562, and Cys672, which are not found in the plasminogen beta-chain. Disulfide bond analysis showed a Cys527-Cys562, but also a Cys588-Cys672. Coopting Cys588 by Cys672 prevented the expected formation of a disulfide between alpha-chain Cys468 and beta-chain Cys588. Concomitant studies determined structures of oligosaccharides at the three Asn-linked glycosylation sites of MSP. The oligosaccharides at the three Asn loci are heterogeneous; 11 different sugars were identified, all being sialylated fucosyl biantennary structures. We also located the pro-MSP signal peptide cleavage site at Gly18-Gln19 and the scissile bond for formation of mature MSP at Arg483-Val484.  相似文献   

13.
This report describes the structure of soluble human stem cell factor isolated from the conditioned medium of Chinese hamster ovary (CHO) cells transfected with stem cell factor (SCF) cDNA, which encodes a leader sequence plus 248 additional amino acids. The 248 amino acids include a hydrophobic transmembrane region at positions 190-212. The isolated material is glycosylated and three bands (apparent M(r) 28,000, M(r) 35,000, and M(r) 40,000) are evident by sodium dodecyl sulfate (SDS)-polyacrylamide gel electrophoresis. After complete deglycosylation, the molecular weight by SDS-polyacrylamide gel electrophoresis is 18,000-19,000. Structural analyses of the intact SCF, the deglycosylated SCF, and a deglycosylated C-terminal peptide were performed by laser desorption, fast atom bombardment, or electrospray mass spectrometry. Pulse-labeling of cells with 35S-labeled Met and Cys resulted in cell-associated glycosylated SCF of M(r) 33,000-45,000 which was converted to M(r) 33,000 by in vitro treatment with glycosidases. During a chase with unlabeled Met and Cys, labeled SCF of M(r) 28,000, M(r) 35,000, and M(r) 40,000 appeared in the medium; it was converted to M(r) 18,000-19,000 by glycosidase treatment. SCF at the surface of the transfected CHO cells could be demonstrated by immunofluorescence. The data obtained indicate that the recombinant human stem cell factor, as isolated, represents proteolytically processed forms containing amino acids 1-165, derived from the initially synthesized membrane-bound form of 248 amino acids. Further characterization indicated that the M(r) 28,000 form is glycosylated at Asn120, the M(r) 35,000 form at Asn120 and Asn65, and the M(r) 40,000 form at Asn120, Asn93, and Asn65. Each form also contains O-linked carbohydrate. The N-linked glycosylation, particularly that at Asn93 and at Asn65, adversely affects in vitro biological activity and receptor binding.  相似文献   

14.
We analysed the oligosaccharides of a human IgM produced bya human-human-mouse hybridoma at each of its five conservedheavy chain glycosylation sites. Consistent with previous reports,this IgM possesses sialylated oligosaccharides at Asn171, Asn332and Asn395, and high-mannose-type oligosaccharides at Asn402.In contrast to previous reports for human IgMs, we find thatAsn563 is not occupied by oligosaccharide on perhaps 25% ofIgM heavy chains, while occupied Asn563 sites contain both high-mannose-typeand sialylated oligosaccharides. These latter results are consistentwith the glycosylation at Asn563 previously reported for themouse MOPC 104E IgM. We demonstrate that both the human hybridomaIgM and the mouse MOPC 104E IgM are mixtures of pentamers andhexamers, raising the possibility that the unique findings concerningthe glycosylation at Asn563 in this study and the previous studyof the MOPC 104E IgM could be related, at least in part, tothe different packing requirements of the hexameric geometryand the accessibility of oligosaccharides in the hexameric geometryfor processing to complex type. In addition, we used high-pHanion-exchange (HPAE) chromatography, neutral anion-exchangechromatography, fluorophore-assisted carbohydrate electrophoresisand Western blots to compare the oligosaccharide compositionsof the human hybridoma IgM, pooled human serum IgM and two mousemonoclonal IgMs (MOPC 104E and TEPC 183). Of note is the presenceof N-glycolylneuraminic acid (NeuGc) and N-acetymeuraminic acid(NeuAc) at a 2:1 ratio in the oligosaccharides of the humanhybridoma IgM. The presence of both NeuGc and NeuAc complicatesthe interpretation of HPAE chromato-graphs. glycosylation high-pH anion-exchange chromatography human IgM human—mouse hybridoma oligosaccharide  相似文献   

15.
This report describes the N-glycosylation site mapping of human serotransferrin (h-STF). Reduced and S-carboxymethylated h-STF was digested with trypsin or chymotrypsin. Glycopeptides in the proteolytic digests were isolated by serial concanavalin A (Con A), Sambucus nigra agglutinin (SNA), and Phaseolus vulgaris leukoagglutinin (LPHA) affinity chromatography and subjected to preliminary analysis by 1H NMR spectroscopy. The glycopeptide fractions were then individually digested with N-glycanase. One part of the digest of each fraction was analyzed by fast atom bombardment-mass spectrometry (FAB-MS) to identify the peptide sequences of the glycosylation sites. The other part was used to isolate the oligosaccharide by the corresponding lectin affinity chromatography and to characterize the structures of the isolated oligosaccharides by 1H NMR spectroscopy and FAB-MS. The oligosaccharides in the Con A-bound fraction were shown to have bi-alpha(2-->6)-sialyl, diantennary structures. The SNA-bound fraction was shown to contain trisialyl, triantennary structures. Di- and triantennary oligosaccharides were found to occur on each of the two N-glycosylation sites of h-STF (Asn413 and Asn611) in the ratio of approximately 85:15. The SNA-bound glycopeptides were further fractionated by LPHA affinity chromatography. Two different oligosaccharides were characterized, namely, a trisialyl 2,4-triantennary and a trisialyl 2,6-triantennary glycan. The ratio of 2,4-triantennary vs 2,6-triantennary oligosaccharides attached to glycosylation site Asn413 was found to be approximately 5:1, whereas the two isomeric triantennary oligosaccharides were found to be attached to glycosylation site Asn611 in the ratio approximately 1:1.  相似文献   

16.
Human calcitonin receptor (hCTR) subtypes contain three or four potential Asn-linked glycosylation sites in their extracellular amino termini. The role of glycosylation in hCTR function has not been identified, but it has been suggested that inhibition of glycosylation does not affect binding or signaling. To determine the role of glycosylation in hCTR biology, we studied the effects of inhibition of glycosylation and of substitution of Asn residues that are potential glycosylation sites. Native and mutated hCTRs were studied after transient expression in monkey kidney COS-1 cells. Tunicamycin, administered as part of a treatment protocol that inhibited glycosylation of all expressed receptors, decreased salmon calcitonin (sCT) binding affinities and signaling potencies at hCTRs with three or four potential glycosylation sites. In hCTR3, which contains three potential glycosylation sites at positions 26, 78, and 83, site-specific substitution of Asn-26 by Ala had no effect on sCT binding affinity or potency, whereas substitution of Asn-78 or Asn-83 lowered sCT affinity and potency. A mutant hCTR3 in which all three Asn residues were substituted with Ala exhibited no high-affinity sCT binding and potencies of several calcitonin analogues that were more than 100-fold lower than that of native hCTR3. Our data show that glycosylation is important for high-affinity binding and potency of calcitonin analogues at hCTRs.  相似文献   

17.
The primary structure of a soluble form of the CD4 receptor (sCD4) expressed in Chinese hamster ovary cells has been confirmed by mass spectrometric peptide mapping and and tandem mass spectrometry. These studies corroborated 95% of the 369-amino acid-long sequence and established the fidelity of translation of the NH2 and COOH terminal including the absence of "ragged ends." The arrangement of the three disulfide bonds in recombinant sCD4 was also established by mass spectrometry and comparative high performance liquid chromatography mapping and shown to be identical to that expected from previous studies of intrachain disulfide bonding in T4 antigens derived from sheep and mouse. No other arrangements of disulfides were detected. Carbohydrate mapping by mass spectrometry was used to establish that both potential Asn-linked glycosylation sites in sCD4 (Asn271 and Asn300) have oligosaccharides attached. Structural characterization by mass spectrometry and methylation analysis of the heterogeneous family of oligosaccharides at each of the specific attachment sites indicates that the major components of both families of oligosaccharides have the following biantennary structures: (Formula, see text) where m + n = 0-2, and x = 0,1. Minor carbohydrate components having three N-acetylneuraminic acid (NeuAc) groups and an additional hexose-hexosamine unit were detected by high performance anion-exchange chromatography.  相似文献   

18.
ADP-ribosylating immunotoxins are generally expressed in Escherichia coli and then refolded in vitro. Because the efficiency of the in vitro refolding process decreases with the number of protein domains and internal disulfide bonds, these immunotoxins have been generally limited to single-chain monovalent structures. We now show that using the hamster cell line CHO K1 RE1.22c (J. M. Moehring and T. J. Moehring, 1979, Somat. Cell Genet. 5, 453-468) that has been mutated to ADP-ribosylation insensitivity, a level of 4 microg/ml of a truncated anti-T cell immunotoxin, DT390-scFvUCHT1, can be secreted into the medium. This immunotoxin is glycosylated at the two potential N-linked glycosylation sites in the toxin moiety: positions 16-18 in the A chain and residues 235-237 in the B chain. The glycosylated immunotoxin is relatively nontoxic (IC(50) 4.8 x 10(-10) M). Removal of the N-linked oligosaccharides by N-glycosidase F treatment or mutations at the two N-linked glycosylation sites results in a highly active immunotoxin with an IC(50) of 4 x 10(-12) M toward CD3(+) Jurkat cells. This is a 12-fold increase in toxicity over the same immunotoxin harvested from E. coli periplasm without refolding. A single Asn(235) Ala mutation that removed the B chain glycosylation was nearly as toxic as the double mutant. This suggests that B chain glycosylation is the major cause for the loss of toxicity.  相似文献   

19.
Follistatin (FS), a glycoprotein, plays an important role in cell growth and differentiation through the neutralization of the biological activities of activins. In this study, we analyzed the glycosylation of recombinant human FS (rhFS) produced in Chinese hamster ovary cells. The results of SDS-PAGE and MALDI-TOF MS revealed the presence of both non-glycosylated and glycosylated forms. FS contains two potential N-glycosylation sites, Asn95 and Asn259. Using mass spectrometric peptide/glycopeptide mapping and precursor-ion scanning, we found that both N-glycosylation sites were partially glycosylated. Monosaccharide composition analyses suggested the linkages of fucosylated bi- and triantennary complex-type oligosaccharides on rhFS. This finding was supported by mass spectrometric oligosaccharide profiling, in which the m/z values and elution times of some of the oligosaccharides from rhFS were in good agreement with those of standard oligosaccharides. Site-specific glycosylation was deduced on the basis of the mass spectra of the glycopeptides. It was suggested that biantennary oligosaccharides are major oligosaccharides located at both Asn95 and Asn259, whereas the triantennary structures are present mainly at Asn95.  相似文献   

20.
The transforming growth factor-beta 1 (TGF beta 1) and -beta 2 (414) precursors both contain three predicted sites of N-linked glycosylation within their pro regions. These are located at amino acid residues 72, 140, and 241 for the TGF beta 2 (414) precursor and at residues 82, 136, and 176 for the TGF beta 1 precursor; both proteins contain mannose-6-phosphate (M-6-P) residues. The major sites of M-6-P addition are at Asn (82) and Asn (136), the first two sites of glycosylation, for the TGF beta 1 precursor. We now show that the major site of M-6-P addition within the TGF beta 2 (414) precursor is at Asn241, the third glycosylation site. To determine the importance of N-linked glycosylation to the secretion of TGF beta 1 and -beta 2, site-directed mutagenesis was used to change the Asn residues to Ser residues; the resulting DNAs were transfected into COS cells, and their supernatants were assayed for TGF beta activity. Substitution of Asn (241) of the TGF beta 2 (414) precursor resulted in an 82% decrease in secreted TGF beta 2 bioactivity. Mutation at Asn72 resulted in a 44% decrease, while mutation at Asn140 was without effect. Elimination of all three glycosylation sites resulted in undetectable levels of TGF beta 2. These results were compared with similar mutations made in the cDNA encoding the TGF beta 1 precursor. Mutagenesis of the two M-6-P-containing sites (Asn82 and Asn136) resulted in an 83% decrease in secreted TGF beta 1; replacement of Asn82 and Asn136 with Ser individually resulted in 85% and 42% decreases in activity, respectively. Substitution of Asn176 with Ser was without effect, while substitution of all three sites of glycosylation resulted in undetectable levels of TGF beta 1 activity, similar to the results obtained with TGF beta 2. The nine Cys residues within the mature region of TGF beta 1 were mutated to serine, and their effects on TGF beta 1 secretion were evaluated. Mutation of most Cys residues resulted in undetectable levels of TGF beta 1 protein or activity in conditioned medium. Mutation of Cys (355) led to the secretion of inactive TGF beta 1 monomers, suggesting that this residue is either directly involved in dimer formation or required for correct interchain disulfide bond formation.  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