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.
Eosinophil granule major basic protein 2 (MBP2 or major basic protein homolog) is a paralog of major basic protein (MBP1) and, similar to MBP1, is cytotoxic and cytostimulatory in vitro. MBP2, a small protein of 13,433 Da molecular weight, contains 10 cysteine residues. Mass spectrometry shows two cystine disulfide linkages (Cys20–Cys115 and Cys92–Cys107) and 6 cysteine residues with free sulfhydryl groups (Cys2, Cys23, Cys42, Cys43, Cys68, and Cys96). MBP2, similar to MBP1, has conserved motifs in common with C-type lectins. The disulfide bond locations are conserved among human MBP1, MBP2 and C-type lectins.  相似文献   

2.
Pathways of oxidative folding of disulfide proteins display a high degree of diversity and vary among two extreme models. The BPTI model is defined by limited species of folding intermediates adopting mainly native disulfide bonds. The hirudin model is characterized by highly heterogeneous folding intermediates containing mostly non-native disulfide bonds. αLA-IIIA is a 3-disulfide variant of α-lactalbumin (αLA) with a 3-D conformation essentially identical to that of intact αLA. αLA-IIIA contains 3 native disulfide bonds of αLA, two of them are located at the calcium binding β-subdomain (Cys61–Cys77 and Cys73–Cys91) and the third bridge is located within the α-helical domain of the molecule (Cys28–Cys111). We investigate here the pathway of oxidative folding of fully reduced αLA-IIIA with and without stabilization of its β-subdomain by calcium binding. In the absence of calcium, the folding pathway of αLA-IIIA was shown to resemble that of hirudin model. Upon stabilization of β-sheet domain by calcium binding, the folding pathway of αLA-IIIA exhibits a striking similarity to that of BPTI model. Three predominant folding intermediates of αLA-IIIA containing exclusively native disulfide bonds were isolated and structurally characterized. Our results further demonstrate that stabilization of subdomains in a protein may dictate its folding pathway and represent a major cause for the existing diversity in the folding pathways of the disulfide-containing proteins.  相似文献   

3.
A full-length cDNA coding lipoprotein lipase (LPL) was cloned from liver of adult common carp (Cyprinus carpio Var. Jian) by RT-PCR and rapid amplification of cDNA ends (RACE) approaches. The cDNA obtained was 2,411 bp long with a 1,524 bp open reading frame (ORF) encoding 507 amino acids. This amino acid sequence contains two structural regions: N-terminus (24–354 residues) and C-terminus (355–507 residues). Before N-terminus, 1–23 residues is signal peptide, 6–23 residues is transmembrance helix. At N-terminus, some conversed functional sites were found, including two N-linked glycosylation sites Asn41 and Asn88; one catalytic triad Ser174, Asp198 and His283; one conserved heparin-binding site Arg321 to Arg324 (RKNR); eight cysteines residues Cys69 and Cys82, Cys258 and Cys281, Cys306 and Cys325, Cys317 and Cys320 which are involved in four disulfide bridges; one polypeptide “lid” that participates in substrate specificity. At C-terminus, Asn401 is another N-linked glycosylation site, and Trp434 and Trp435 (WW) is lipid-binding site. The amino acid sequence has a high similarity, and shows similar structural features to LPL of other species. Tissue distribution of LPL mRNA in liver, head kidney, mesenteric adipose tissue, heart and white muscle of common carp was analyzed by semi-quantitative RT-PCR method using β-actin gene as internal control. The result showed that the expressions of LPL mRNA were detected in all examined tissues of common carp. The expression levels of LPL in the mesenteric adipose tissue was highest among these tissues, following in liver and head kidney, and the lowest expression was found in heart and white muscle.  相似文献   

4.
VP1 is the major coat protein of murine polyomavirus and forms virus-like particles (VLPs) in vitro. VLPs consist of 72 pentameric VP1 subunits held together by a terminal clamp structure that is further stabilized by disulfide bonds and chelation of calcium ions. Yeast-derived VLPs (yVLPs) assemble intracellularly in vivo during recombinant protein production. These in vivo assembled yVLPs differ in several properties from VLPs assembled in vitro from bacterially produced pentamers. We found several intermolecular disulfide linkages in yVLPs involving 5 of the 6 cysteines of VP1 (Cys115–Cys20, Cys12–Cys20, Cys16–Cys16, Cys12/ Cys16–Cys115, and Cys274–Cys274), indicating a highly coordinated disulfide network within the in vivo assembled particles involving the N-terminal region of VP1. Cryoelectron microscopy revealed structured termini not resolved in the published crystal structure of the bacterially expressed VLP that appear to clamp the pentameric subunits together. These structural features are probably the reason for the observed higher stability of in vivo assembled yVLPs compared with in vitro assembled bacterially expressed VLPs as monitored by increased thermal stability, higher resistance to trypsin cleavage, and a higher activation enthalpy of the disassembly reaction. This high stability is decreased following disassembly of yVLPs and subsequent in vitro reassembly, suggesting a role for cellular components in optimal assembly.  相似文献   

5.
A fast and efficient microwave-assisted solid phase peptide synthesis (MW-SPPS) of a 51mer peptide, the main heparin-binding site (60–110) of human pleiotrophin (hPTN), using 2-chlorotrityl chloride resin (CLTR-Cl) following the 9-fluorenylmethyloxycarbonyl/tert-butyl (Fmoc/tBu) methodology and with the standard N,N′-diisopropylcarbodiimide/1-hydroxybenzotriazole (DIC/HOBt) coupling reagents, is described. An MW-SPPS protocol was for the first time successfully applied to the acid labile CLTR-Cl for the faster synthesis of long peptides (51mer peptide) and with an enhanced purity in comparison to conventional SPPS protocols. The synthesis of such long peptides is not trivial and it is generally achieved by recombinant techniques. The desired linear peptide was obtained in only 30 h of total processing time and in 51% crude yield, in which 60% was the purified product obtained with 99.4% purity. The synthesized peptide was purified by reversed phase high performance liquid chromatography (RP-HPLC) and identified by electrospray ionization mass spectrometry (ESI-MS). Then, the regioselective formation of the two disulfide bridges of hPTN 60–110 was successfully achieved by a two-step procedure, involving an oxidative folding step in dimethylsulfoxide (DMSO) to form the Cys77–Cys109 bond, followed by iodine oxidation to form the Cys67–Cys99 bond.  相似文献   

6.
The oligomeric state of human porphobilinogen synthase (PBGS) [EC.4.2.1.24] is homooctamer, which consists of conformationally heterogenous subunits in the tertiary structure under air-saturated conditions. When PBGS is activated by reducing agent with zinc ion, a reservoir zinc ion coordinated by Cys223 is transferred in the active center to be coordinated by Cys122, Cys124, and Cys132 (Sawada et al. in J Biol Inorg Chem 10:199–207, 2005). The latter zinc ion serves as an electrophilic catalysis. In this study, we investigated a conformational change associated with the PBGS activation by reducing agent and zinc ion using analytical ultracentrifugation, negative staining electron microscopy, native PAGE, and enzyme activity staining. The results are in good agreement with our notion that the main component of PBGS is octamer with a few percent of hexamer and that the octamer changes spatial subunit arrangement upon reduction and further addition of zinc ion, accompanying decrease in f/f 0. It is concluded that redox-regulated PBGS activation via cleavage of disulfide bonds among Cys122, Cys124, and Cys132 and coordination with zinc ion is closely linked to change in the oligomeric state.  相似文献   

7.
Summary α-conotoxin EI is an 18-residue peptide (RDOCCYHPTCNMSNPQIC; 4–10, 5–18) isolated from the venom ofConus ermineus, the only fish-hunting cone snail of the Atlantic Ocean. This peptide targets specifically the nicotinic acetylcholine receptor (nAChR) found in mammalian skeletal muscle and the electric organTorpedo, showing a novel selectivity profile when compared to other α-conotoxins. The 3D structure of EI has been determined by 2D-NMR methods in combination with dynamical simulated annealing protocols. A total of 133 NOE-derived distances were used to produce 13 structures with minimum energy that complied with the NOE restraints. The structure of EI is characterized by a helical loop between THr9 and Met12 that is stabilized by the Cys4-Cys10 disulfide bond and turns involving Cys4-Cys5 and Asn14-Pro15. Other regions of the peptide appear to be flexible. The overall fold of EI is similar to that of other α4/7-conotoxins (PnIA/B, MII, EpI). However, unlike these other α4/7-conotoxins, EI targets the muscular type nAChR. The differences in selectivity can be attributed to differences in the surface charge distribution among these α4/7-conotoxins. The implications for binding of EI to the muscular nAChR are discussed with respect to the current NMR structure of EI. Supplementary material available:1H resonance assignments of α-conotoxin EI.  相似文献   

8.
Summary Four enhanced carbonyl carbon resonances were observed whenStreptomyces subtilisin inhibitor (SSI) was labeled by incorporating specifically labeled [1-13C]Cys. The13C signals were assigned by the15N,13C double-labeling method along with site-specific mutagenesis. Changes in the spectrum of the labeled protein ([C]SSI) were induced by reducing the disulfide bonds with various amounts of dithiothreitol (DTT). The results indicate that, in the absence of denaturant, the Cys71-Cys101 disulfide bond of each SSI subunit can be reduced selectively. This disulfide bond, which is in the vicinity of the reactive site scissile bond Met73-Val74, is more accessible to solvent than the other disulfide bond. Cys35-Cys50, which is embedded in the interior of SSI. This half-reduced SSI had 65% of the inhibitory activity of native SSI and maintained a conformation similar to that of the fully oxidized SSI. Reoxidation of the half reduced-folded SSI by air regenerates fully active SSI which is indistinguishable with intact SSI by NMR. In the presence of 3 M guanidine hydrochloride (GuHCl), however, both disulfide bonds of each SSI subunit were readily reduced by DTT. The fully reduced-unfolded SSI spontaneously refolded into a native-like structure (fully reduced-folded state), as evidenced by the Cys carbonyl carbon chemical shifts, upon removing GuHCl and DTT from the reaction mixture. The time course of disulfide bond regeneration from this state by air oxidation was monitored by following the NMR spectral changes and the results indicated that the disulfide bond between Cys71 and Cys101 regenerates at a much faster rate than that between Cys35 and Cys50.Nomenclature of the various states of SSI that are observed in the present study Fully oxidized-folded native or intact (without GuHCl or DTT) - half reduced-folded (Cys71-Cys101 reduced; DTT without GuHCl) - inversely half reduced-folded (Cys35-Cys50 reduced; a reoxidation intermediate from fully reduced-folded state) - fully reduced-unfolded (reduced by DTT in the presence of GuHCl) - fully reduced-folded (an intermediate state obtained by removing DTT and GuHCl from the fully reduced-unfolded SSI reaction mixture)  相似文献   

9.
Chemical synthesis of kurtoxin, a T-type calcium channel blocker   总被引:1,自引:0,他引:1  
Kurtoxin isolated from the venom of scorpion, Parabuthus transvaalicus, is a 63-residue peptide with four intramolecular disulfide bonds which inhibits low-threshold T-type Ca2+channels. Kurtoxin was synthesized by native chemical ligation involving the coupling of (1--26)-thioester peptide and Cys27-(28--63)-peptide. The former was synthesized by standard solid-phase peptide synthesis (SPPS) with Boc chemistry, while the latter was sequentially assembled from three protected segments onto a resin-bound C-terminal segment in a chloroform--phenol mixed solvent followed by deprotection reaction using HF. Each protected segment used for the coupling on a solid support was prepared on an N-[9-(hydroxymethyl)-2-fluorenyl] succinamic acid (HMFS) resin and detached from the resin by treatment with 20% Et 3N in DMF to produce it in the form of an α-carboxylic acid. Synthetic kurtoxin obtained after the oxidative folding reaction was found to be identical with the natural product by means of several analytical procedures, and its disulfide structure was determined for the first time to be Cys12-Cys61, Cys16-Cys37, Cys23-Cys44 and Cys27-Cys46 by peptide mapping, sequence analysis and mass measurements.  相似文献   

10.
Human neurotrophin-3 (NT-3) is a member of the nerve growth factor (NGF) family of neurotrophic factors, and the recombinant protein is being developed as a therapeutic for neurodegenerative diseases. The final product purity and lot-to-lot variation are monitored routinely by peptide mapping. However, only the N-terminal region of NT-3 was susceptible to proteolysis under native conditions. Complete digestion required that the protein be chemically modified by reduction and S-alkylation prior to proteolysis. Complete proteolytic degradation of the protein was achieved simply by an intial denaturation of NT-3 in 6 M guanidinium chloride (pH 6) for 2 hr at 37°C, followed by a tenfold dilution with the digestion buffer (0.1 M Tris-HCl, 1 mM CaCl2 at pH 7.0) and immediate addition of chymotrypsin at 1% by weight. Direct comparison of the peptide map with an identical aliquot that had been reduced and alkylated also allowed the establishment of the cystine linkages present in NT-3: Cys14 to Cys79, Cys57 to Cys108, and Cys67 to Cys110. This disulfide structure is homologous to the NGF family of neurotrophic factors.  相似文献   

11.
The crucial step of folding of recombinant proteins presents serious challenges to obtaining the native structure. This problem is exemplified by insulin‐like growth factor (IGF)‐I which when refolded in vitro produces the native three‐disulfide structure, an alternative structure with mispaired disulfide bonds and other isomeric forms. To investigate this phenomenon we have examined the refolding properties of an analog of IGF‐I which contains a 13‐amino acid N‐terminal extension and a charge mutation at position 3 (Long‐ [Arg3]IGF‐I). Unlike IGF‐I, which yields 45% of the native structure and 24% of the alternative structure when refolded in vitro, Long‐[Arg3]GF‐I yields 85% and 10% of these respective forms. To investigate the interactions that affect the refolding of Long‐[Arg3]IGF‐I and IGF‐I, we acid‐trapped folding intermediates and products for inclusion in a kinetic analysis of refolding. In addition to non‐native intermediates, three native‐like intermediates were identified, that appear to have a major role in the in vitro refolding pathway of Long‐[Arg3]IGF‐I; a single‐disulfide Cys18–Cys61 intermediate, an intermediate with Cys18–Cys61 and Cys6–Cys48 disulfide bonds and another with Cys18–Cys61 and Cys47–Cys52 disulfide bonds. Furthermore, from our kinetic analysis we propose that the Cys18‐Cys61, Cys6‐Cys48 intermediate forms the native structure, not by the direct formation of the last (Cys47‐Cys52) disulfide bond, but by rearrangement via the Cys18–Cys61 intermediate and a productive Cys18–Cys61, Cys47–Cys52 intermediate. In this pathway, the last disulfide bond to form involves Cys6 and Cys48. Finally, we apply this pathway to IGF‐I and conclude that the divergence in the in vitro folding pathway of IGF‐I is caused by non‐native interactions involving Glu3 that stabilize the alternative structure. © 1999 John Wiley & Sons, Inc. Biotechnol Bioeng 62: 693–703, 1999.  相似文献   

12.
Summary ShK toxin, a 35-residue peptide isolated from the Caribbean sea anemone Stichodactyla helianthus, is a potent inhibitor of the Kv 1.3 potassium channel in lymphocytes. The natural toxin contains three disulfide bonds. The disulfide pairings of the synthetic ShK toxin were elucidated as a prerequisite for studies on its structure-function relationships. The toxin was fragmented at pH 6.5 using either thermolysin or a mixture of trypsin and chymotrypsin followed by thermolysin. The fragments were isolated by RP-HPLC and were identified by sequence analysis and MALDI-TOF mass spectrometry. The three disulfides were unambiguously identified in either proteolytic digest: Cys3 to Cys35, Cys12 to Cys28 and Cys17 to Cys32. The Cys3-Cys35 disulfide, linking the amino- and carboxyl-termini, defines the characteristic cyclic structure of the molecule. A similar disulfide pairing motif is found in the snake venom-derived potassium channel blocker dendrotoxin and the mammalian antibiotic peptide defensins.  相似文献   

13.
The conformational stability of human epidermal growth factor (EGF) and the structure of denatured EGF were investigated using the technique of disulfide scrambling. Under denaturing conditions and in the presence of a thiol catalyst, the native EGF denatures by shuffling its three native disulfide bonds and converts to a mixture of scrambled isomers. Analysis by HPLC reveals that the denatured EGF is composed of about 10 fractions of scrambled isomers. The heterogeneity varies under different denaturing conditions, with the heat-denatured samples exhibiting the highest degree of heterogeneity. The disulfide structures of eight major scrambled isomers of EGF were determined. The most predominant isomer adopts the bead-form structure with disulfide bonds bridged by three pairs of neighboring cysteines: Cys6-Cys14, Cys20-Cys31, and Cys33-Cys42. The denaturation curve of EGF is determined by the relative yield of the scrambled and native species of EGF. EGF is a highly stable molecule and can be effectively denatured only by guanidine chloride at a concentration of greater than 4–5 M. At 8 M urea, less than 16% of the native EGF was denatured. The unusual conformational stability of EGF was compared with that of eight different disulfide proteins that were similarly characterized by the method of disulfide scrambling.  相似文献   

14.
The interactions of theω-amino acid ligandsε-aminocaproic acid andp-benzylaminesulphonic acid with the isolated kringle 4 domain from human plasminogen have been investigated by1H-nuclear magnetic resonance spectroscopy at 300 and 600 MHz. Overall, the data indicate that binding either ligand does not cause the kringle to undergo significant conformational changes. When p-benzylaminesulphonic acid is in excess relative to the kringles, progressive exchange-broadening and high field chemical shifts are observed for the proton resonances of the ligand. The largest effect is seen at the amino end of the molecule, which indicates that the — NH 3 + group of the ligand penetrates deeper into the binding site than does the — SO 3 - . Ligand-binding causes signals from the ring-current shifted Leu46 CH 3 δ .δ groups and from a number of aromatic side-chains to shift. Depending on the ligand, the latter include Tyr-II (Tyr50), Tyr-V (an immobile ring), His-II and His-III imidazole groups and the three Trp indole groups present in kringle 4. In particular,p-benzylaminesulphonic acid-binding induces large high field shifts on the Trp-II H6 triplet and the Trp-III (Trp72) H2 singlet. On the other hand,ε-aminocaproic acid bound to kringle 4 exhibits large chemical shifts of its CH2 proton resonances, which indicates that the lysine-binding site is rich in aromatic side chains. Overhauser experiments centered on thep-benzylaminesulphonic acid H2,6 and H3,5 aromatic transitions as well as on the shifted Trp-II and Trp-III signals reveal efficient cross-relaxation between these two indole side chains and thep-benzylaminesulphonic acid ring. These experiments also show that the side chains from Phe64, Tyr-II (Tyr50), Tyr-IV, and His-II (His31) interact with the ligand. In combination with reported chemical modification experiments that show requirement of Asp57, Arg71 and Trp72 integrity for ligand-binding, our study underscores the relevance of the Cys51-Cys75 loop in defining the kringles’ lysine-binding site. Furthermore, the Cys22-Cys63 loop is folded so as to place His31, His33, Tyr41 and Leu46 in proximity to the binding site. The involvement of residues within the Cys51-Cys75 loop in ligand-binding suggests that Trp-II and Tyr-IV may correspond to Trp62 and Tyr74, respectively. As shown by Overhauser experiments, these two residues are in close contact with each other. From these studies and from the shielding and deshielding effects caused byp-benzylaminesulphonic acid, we suggest that the ligand is sandwiched between the indole rings of Trp-II and Trp-III, which form part of the hydrophobic binding site.  相似文献   

15.
We report the nearly complete 1H, 13C and 15N resonance assignments of the oxidized form (Cys67–Cys70) of the N-terminal domain of PilB from Neisseria meningitidis. Secondary structure determination using CSI method and TALOS leads mainly to the prediction of 7 α-helical and 5 β-sheet parts.  相似文献   

16.
A series of N-substituted maleimides was shown to effectively inactivate bremazocine binding to δ opioid receptors. Apparent second order rate constants for inactivation increased with increasing size of the N-substituent: N-methyl < N-ethyl < N-butyl < N-phenylmaleimide. It is suggested that the positive chain length effect is attributed to nonpolar interactions with the receptor in the vicinity of the reactive group. Binding to μ and δ opioid receptors was equally sensitive to inactivation by (2-aminoethyl)methanethiosulfonate; the [2-(trimethylammonium)ethyl] and (2-sulfonatoethyl) derivatives were less active. Site-directed mutagenesis of the μ opioid receptor indicated that Cys159, Cys190, Cys235, Cys292, or Cys321, residing in transmembrane domain 3, 4, 5, 6, and 7, respectively, werenot the site of modification. Special issue dedicated to Dr. Eric J. Simon.  相似文献   

17.
Hen ovalbumin contains one cystine disulfide (Cys73-Cys120) and four cysteine sulfhydryl groups (Cys11,Cys30,Cys367, and Cys382) in a single polypeptide chain of 385 amino acid residues. To investigate whether or not such a structure is shared by related avian species, the contents of disulfide-involved half-cystine residues and their positions in the primary structure of ovalbumins from five species were compared with those of hen ovalbumin. Ovalbumins were alkylated with a fluorescent dye, IAEDANS, under disulfide-reduced and disulfide-intact conditions and digested with a number of proteolytic enzymes. The sequences were deduced from peptides containing half-cystine residues labeled with the fluorescent dye. The results showed that the number of free cysteine sulfhydryl groups of ovalbumins was different among the species, three for guinea fowl and turkey (Cys11, Cys367, and Cys382); and two for Pekin duck, mallard duck, and Emden goose (Cys11 and Cys331). On the other hand, a single intrachain disulfide bond could be identified from ovalbumins of five species using a combination of peptide mapping and N-terminal amino acid sequencing analysis under reduced and non-reduced conditions, in which the intrachain disulfide bond was like that of hen ovalbumin (Cys73-Cys120). The results also indicated that the variations in amino acid sequences on these peptides containing half-cystine residues bear a close relationship with the phylogeny of the six species.  相似文献   

18.
Transglutaminase 2 (TG2) in the extracellular matrix is largely inactive but is transiently activated upon certain types of inflammation and cell injury. The enzymatic activity of extracellular TG2 thus appears to be tightly regulated. As TG2 is known to be sensitive to changes in the redox environment, inactivation through oxidation presents a plausible mechanism. Using mass spectrometry, we have identified a redox-sensitive cysteine triad consisting of Cys230, Cys370, and Cys371 that is involved in oxidative inactivation of TG2. Within this triad, Cys370 was found to participate in disulfide bonds with both Cys230 and its neighbor, Cys371. Notably, Ca2+ was found to protect against formation of these disulfide bonds. To investigate the role of each cysteine residue, we created alanine mutants and found that Cys230 appears to promote oxidation and inactivation of TG2 by facilitating formation of Cys370–Cys371 through formation of the Cys230–Cys370 disulfide bond. Although vicinal disulfide pairs are found in several transglutaminase isoforms, Cys230 is unique for TG2, suggesting that this residue acts as an isoform-specific redox sensor. Our findings suggest that oxidation is likely to influence the amount of active TG2 present in the extracellular environment.  相似文献   

19.
The positions of the disulfide bonds of huwentoxin-I, a neurotoxin from the spiderSelenocosmia huwena, have been determined. The existence of three disulfide bonds in the native toxin was demonstrated by mass spectroscopy and the lack of reactivity with a thiol reagent. The assignment procedure involved a combination of tryptic digestion of the native toxin and sequence analysis of both intact andin situ S-carboxymethylated toxin.In situ carboxymethylation is shown to be a useful procedure in sequencing of cysteine- and cystine-containing peptides. Sequence analysis of the intact, cross-linked toxin indicated that no amino acid phenylthiohydantoin (PTH) derivative is seen for the first half-cystine in a cross-linked pair, but that the PTH of dehydroalanine, which can be detected at 313 nm, is seen at the position of the second half-cystine. By sequencing disulfide cross-linked tryptic fragments, the three disulfide linkages in huwentoxin-I could be assigned as Cys2-Cys17, Cys9-Cys22, and Cys16-Cys29.  相似文献   

20.
The uncovering enzyme (UCE) removes N-acetylglucosamine from lysosomal enzymes to uncover the mannose 6-phosphate (Man-6-P) determinant necessary for targeting these enzymes to lysosomes. Failure to create the Man-6-P determinant is one cause of lysosomal storage diseases. Despite its medical importance, little structural information about UCE is available. In this report we have developed a model for the membrane proximal portion of the lumenal domain of UCE based on the structure of the EFG-3 and -4 domains of the extracellular segment of the beta chain of integrin V 3. In this model the EGF-like domains of UCE (residues 285–345) are predicted to form a rod-shaped stalk region, similar to the stem region in Golgi glycosyltransferases. This stalk causes the proposed catalytic domain (residues 1–277) to be extended away from the Golgi membrane. A portion of the proposed catalytic domain (residues 85-256) resides in Cluster of Orthologous Group (COG) 4632 with four bacterial proteins but is not homologous to any known eukaryotic proteins. Thus, UCE may have evolved from the fusion of a unique catalytic domain with a common EGF-like stalk domain. We have determined by mass spectrometry that the four disulfide bonds of the proposed catalytic domain are located between Cys2–Cys172, Cys66–Cys99, Cys83–Cys274, and Cys258–Cys265. Finally, we determined that four of the six potential N-linked glycosylation sites are glycosylated (Asn 159, Asn 165, Asn 247, and Asn 317) in COS cells. Published in 2005.  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