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 859 毫秒
1.
Human pancreatic trypsinogens undergo post-translational sulfation on Tyr(154), catalysed by the Golgi-resident enzyme tyrosylprotein sulfotransferase 2. Sequence alignments suggest that the sulfation of Tyr(154) is facilitated by a unique sequence context which is characteristically found in primate trypsinogens. In the search for genetic variants that might alter this sulfation motif, we identified a single nucleotide polymorphism (c.457G>C) in the PRSS2 (serine protease 2, human anionic trypsinogen) gene, which changed Asp(153) to a histidine residue (p.D153H). The p.D153H variant is common in subjects of African origin, with a minor allele frequency of 9.2%, whereas it is absent in subjects of European descent. We demonstrate that Asp(153) is the main determinant of tyrosine sulfation in anionic trypsinogen, as both the natural p.D153H variation and the p.D153N mutation result in a complete loss of trypsinogen sulfation. In contrast, mutation of Asp(156) and Glu(157) only slightly decrease tyrosine sulfation, whereas mutation of Gly(151) and Pro(155) has no effect. With respect to the biological relevance of the p.D153H variant, we found that tyrosine sulfation had no significant effect on the activation of anionic trypsinogen or the catalytic activity and inhibitor sensitivity of anionic trypsin. Taken together with previous studies, the observations of the present study suggest that the primary role of trypsinogen sulfation in humans is to stimulate autoactivation of PRSS1 (serine protease 1, human cationic trypsinogen), whereas the sulfation of anionic trypsinogen is unimportant for normal digestive physiology. As a result, the p.D153H polymorphism which eliminates this modification could become widespread in a healthy population.  相似文献   

2.
Trypsinogen-1 and -2 are well-characterized enzymes that are expressed in the pancreas and also in several other tissues. Many cancers produce trypsinogen isoenzymes that differ from the pancreatic ones with respect to substrate specificity and isoelectric point. These tumor-associated trypsinogens play a pivotal role in cancer progression and metastasis. The differences between these and the pancreatic isoenzymes have been suggested to be caused by post-translational modification, either sulfation or phosphorylation of a tyrosine residue. We aimed to elucidate the cause of these differences. We isolated trypsinogens from pancreatic juice and conditioned medium from a colon carcinoma cell line. Intact proteins, and tryptic and chymotryptic peptides were characterized by electrospray ionization mass spectrometry. We also used immunoblotting with antibody against phosphotyrosine and N-terminal sequencing. The results show that pancreatic trypsinogen-1 and -2 are sulfated at Tyr154, whereas tumor-associated trypsinogen-2 is not. Detachment of a labile sulfogroup could be demonstrated by both in-source dissociation and low-energy collision-induced dissociation in a tandem mass spectrometer. Tyrosine sulfation is an ubiquitous protein modification occurring in the secretory pathway, but its significance is often underestimated due to difficulties in its analysis. Sulfation is an almost irreversible modification that is thought to regulate protein-protein interactions and the activity of proteolytic enzymes. We conclude that the previously known differences in charge, substrate specificity and inhibitor binding between pancreatic and tumor-associated trypsinogens are probably caused by sulfation of Tyr154 in pancreatic trypsinogens.  相似文献   

3.
Human pancreatic secretions contain two major trypsinogen isoforms, cationic and anionic trypsinogen, normally at a ratio of 2 : 1. Pancreatitis, pancreatic cancer and chronic alcoholism lead to a characteristic reversal of the isoform ratio, and anionic trypsinogen becomes the predominant zymogen secreted. To understand the biochemical consequences of these alterations, we recombinantly expressed and purified both human trypsinogens and documented characteristics of autoactivation, autocatalytic degradation and Ca2+-dependence. Even though the two trypsinogens are approximately 90% identical in their primary structure, we found that human anionic trypsinogen and trypsin exhibited a significantly increased (10-20-fold) propensity for autocatalytic degradation, relative to cationic trypsinogen and trypsin. Furthermore, in contrast to the characteristic stimulation of the cationic proenzyme, acidic pH inhibited autoactivation of anionic trypsinogen. In mixtures of cationic and anionic trypsinogen, an increase in the proportion of the anionic proenzyme had no significant effect on the levels of trypsin generated by autoactivation or by enterokinase at pH 8.0 in 1 mm Ca2+- conditions that were characteristic of the pancreatic juice. In contrast, rates of trypsinogen activation were markedly reduced with increasing ratios of anionic trypsinogen under conditions that were typical of potential sites of pathological intra-acinar trypsinogen activation. Thus, at low Ca2+ concentrations at pH 8.0, selective degradation of anionic trypsinogen and trypsin caused diminished trypsin production; while at pH 5.0, inhibition of anionic trypsinogen activation resulted in lower trypsin yields. Taken together, the observations indicate that up-regulation of anionic trypsinogen in pancreatic diseases does not affect physiological trypsinogen activation, but significantly limits trypsin generation under potential pathological conditions.  相似文献   

4.
Isolation and characterization of a cDNA encoding rat cationic trypsinogen   总被引:4,自引:0,他引:4  
A cDNA encoding rat cationic trypsinogen has been isolated by immunoscreening from a rat pancreas cDNA library. The protein encoded by this cDNA is highly basic and contains all of the structural features observed in trypsinogens. The amino acid sequence of rat cationic trypsinogen is 75% and 77% homologous to the two anionic rat trypsinogens. The homology of rat cationic trypsinogen to these anionic trypsinogens is lower than its homology to other mammalian cationic trypsinogens, suggesting that anionic and cationic trypsins probably diverged prior to the divergence of rodents and ungulates. The most unusual feature of this trypsinogen is the presence of an activation peptide containing five aspartic acid residues, in contrast to all other reported trypsinogen activation peptides which contain four acidic amino acid residues. Comparisons of cationic and anionic trypsins reveal that the majority of the charge changes occur in the C-terminal portion of the protein, which forms the substrate binding site. Several regions of conserved charge differences between cationic and anionic trypsins have been identified in this region, which may influence the rate of hydrolysis of protein substrates.  相似文献   

5.
A cDNA clone encoding an anionic form of bovine trypsinogen was isolated from a pancreatic cDNA library. The corresponding 855-nucleotide mRNA contains a short 5' noncoding region of 8 nucleotides and a long 3' noncoding region of 56 nucleotides in addition to a poly(A) tail of at least 50 nucleotides. The deduced amino acid sequence for the anionic pretrypsinogen (247 residues) includes the N-terminal 15-amino-acid signal peptide followed by an 8-amino-acid activation peptide. The zymogen (232 residues) contains an additional C-terminal serine, compared with the amino acid sequence of bovine cationic trypsinogen. The identity between the anionic and cationic forms of bovine trypsinogen (65%) is lower than that existing between the anionic protein and other mammalian anionic trypsinogens (73-85%), suggesting that trypsin gene duplication in mammals occurred prior to the evolutionary events responsible for the species divergence. Bovine pancreatic anionic trypsin possesses all the key amino acids characteristic of the serine protease family.  相似文献   

6.
Anionic and cationic canine trypsinogens were purified from pancreatic juice by affinity chromatography with Trasylol coupled to Sepharose 4B followed by ion exchange chromatography with SP-Sephadex C-50. Automatic N-terminal amino acid sequence determination showed the following structures for the activation peptides: Thr-Pro-Thr-Asp-Asp-Asp-Asp-Lys for anionic trypsinogen, and Phe-Pro-Ile-Asp-Asp-Asp-Asp-Lys for cationic trypsinogen.  相似文献   

7.
Human cationic and anionic trypsins are sulfated on Tyr154, a residue which helps to shape the prime side substrate-binding subsites. Here, we used phage display technology to assess the significance of tyrosine sulfation for the specificity of human trypsins. The prime side residues P1′–P4′ in the binding loop of bovine pancreatic trypsin inhibitor (BPTI) were fully randomized and tight binding inhibitor phages were selected against non-sulfated and sulfated human cationic trypsin. The selection pattern for the two targets differed mostly at the P2′ position, where variants selected against non-sulfated trypsin contained primarily aliphatic residues (Leu, Ile, Met), while variants selected against sulfated trypsin were enriched also for Arg. BPTI variants carrying Arg, Lys, Ile, Leu or Ala at the P2′ position of the binding loop were purified and equilibrium dissociation constants were determined against non-sulfated and sulfated cationic and anionic human trypsins. BPTI variants harboring apolar residues at P2′ exhibited 3–12-fold lower affinity to sulfated trypsin relative to the non-sulfated enzyme, whereas BPTI variants containing basic residues at P2′ had comparable affinity to both trypsin forms. Taken together, the observations demonstrate that the tyrosyl sulfate in human trypsins interacts with the P2′ position of the substrate-like inhibitor and this modification increases P2′ selectivity towards basic side chains.  相似文献   

8.
The activation of human trypsinogens 1 and 2 by porcine enterokinase at pH 5.6 shows that the two human zymogens are equivalent substrates for this enzyme and that both proteins are activated faster than the cationic bovine trypsinogen. At pH 8.0 and in the presence of 20 mM calcium the two human trypsinogens are activated by either human trypsin at the same rate but the affinity of both trypsins is higher for trypsinogen 1 than for trypsinogen 2. Two Ca2+ binding sites are identified in the two human zymogens and their pK(Ca2+) values determined. For trypsinogen 1 the values are respectively of 2.8 and 3.3 for the primary and secondary Ca2+ binding sites, and for trypsinogen 2 of 3.4 and 2.7. These values are markedly different from those obtained for bovine cationic trypsinogen, especially in the case of trypsinogen 1. These results point out a different degree of saturation of the calcium binding sites of the 2 human zymogens that must exist in physiological conditions, suggesting different biological activities of the two trypsinogens.  相似文献   

9.
The two human trypsinogens have been isolated from human pancreatic juice in a sufficient amount to study molecular and structural properties. The purification procedure included filtration on Sephadex G-100 followed by ion-exchange chromatography on DEAE-cellulose. The two trypsinogens represent 19% of total proteins of pancreatic juice. Trypsinogen 1, the major form, is present in a quantity twice that of trypsinogen 2, which is the most anionic protein in human pancreatic juice. The two proteins have partial immunological identity, close molecular weights (23 438 and 25 006 for trypsinogens 1 and 2, respectively) and similar amino acid compositions. The N-terminal sequences are the same for the first 9 residues: Ala-Pro-Phe-Asp4-Lys-Ile. The two proteins differ in the activation peptides released during the transformation to trypsins. Trypsinogen 2 liberates one octapeptide Ala-Pro-Phe-Asp4-Lys while trypsinogen 1 liberates two peptides, the same octapeptide and the pentapeptide (Asp)4-Lys.  相似文献   

10.
The formation of complexes between human trypsinogens and the basic pancreatic trypsin inhibitor is demonstrated by using affinity chromatography on Sepharose coupled to basic pancreatic trypsin inhibitor. This interaction indicates the pre-existence of the active site in human trypsinogens. This active site induces the proteolytic activity of the two zymogens which activate spontaneously at pH 5.6 and pH 8.0 before and after affinity chromatography. The effect of affinity-chromatography on trypsinogen spontaneous activation is not the same on trypsinogens 1 and 2. A striking difference appears between the activation of the two trypsinogens. In all cases, trypsinogen 1 autoactivates more rapidly than trypsinogen 2, except at pH 5.6 in the presence of 10 mM Ca2+, which inhibits the autoactivation of trypsinogen 1. The effect of inherent proteolytic activity of human trypsinogens is discussed in relation to pathological conditions of enterokinase deficiency and acute pancreatitis.  相似文献   

11.
The activation of canine anionic and cationic trypsinogen by enterokinase, trypsin, thrombin, plasmin and extracts from canine granulocytes were studied in vitro. Enterokinase activates both trypsinogens about 1000 times faster than trypsin. The enterokinase-catalyzed activation is not inhibited by the main serum protease inhibitors, alpha-macroglobulin and alpha 1-antitrypsin. alpha-Macroglobulin cannot inhibit the activation of the trypsinogens by trypsin but this reaction is completely inhibited by alpha 1-antitrypsin. The results are discussed in relation to the pathogenesis of acute pancreatitis.  相似文献   

12.
Rat P23 is an isoform of trypsin (ogens) synthesized by rat acinar cells. Expression of P23 is stimulated strongly by caerulein, an analogue of cholecystokinin (CCK). However, the physiological relevance of rat P23 in healthy and pathological conditions such as caerulein-induced pancreatitis is largely unknown. Using recombinant P23 trypsinogen and reconstitution analysis of zymogen autoactivation, unique inhibitor-resistance characteristics of P23 were elucidated. P23 cDNA was expressed in Escherichia coli periplasm, yielding recombinant P23 trypsinogen. Autoactivation of zymogen granule contents from caerulein-induced rat pancreas was also studied. Activation kinetics of P23 by enterokinase was similar to those of rat anionic trypsinogen, which is a major isoform of trypsinogen. Interestingly, rat pancreatic secretory trypsin inhibitor (PSTI), which protects against deleterious activation of trypsinogens in zymogen granules, failed to inhibit P23 trypsin even with four-fold molar excess, at which concentration it effectively inhibited rat anionic trypsin to almost 100%. P23 trypsin also showed marked resistance to proteinaceous trypsin inhibitors such as soybean trypsin inhibitor and aprotinin. P23 trypsin activated by enterokinase dramatically accelerated the cascade of autoactivation of anionic trypsinogen even in the presence of PSTI. Taken together with a previous observation that P23 is specifically upregulated 14-fold by 24-h caerulein infusion, these results suggest that elevated levels of P23 should be taken into consideration in the mechanism of trypsinogens within the pancreas in pathological conditions.  相似文献   

13.
Human S-protein (vitronectin) and hemopexin, two structurally related plasma proteins of similar molecular mass and abundance, were analyzed for tyrosine sulfation. Both proteins were synthesized and secreted by the human hepatoma-derived cell line Hep G2, as shown by immunoprecipitation from the culture medium of [35S]methionine-labelled cells. When Hep G2 cells were labelled with [35S]sulfate, S-protein, but not hemopexin, was found to be sulfated. Half of the [35S]sulfate incorporated into S-protein was recovered as tyrosine sulfate. The stoichiometry of tyrosine sulfation was approximately two mol tyrosine sulfate/mol S-protein. Examination of the S-protein sequence for the presence of the known consensus features for tyrosine sulfation revealed three potential sulfation sites at positions 56, 59 and 401. Tyrosine 56 is the most probable site for stoichiometric sulfation, followed by tyrosine 59 which appears more likely to become sulfated than tyrosine 401. Tyrosines 56 and 59 are located in the anionic region of S-protein which has no homologous counterpart in hemopexin. We discuss the possibility that tyrosine sulfation of the anionic region of S-protein may stabilize the conformation of S-protein in the absence of thrombin-antithrombin III complexes and may play a role in its binding to thrombin-antithrombin III complexes during coagulation.  相似文献   

14.
J Mikkelsen  J Thomsen  M Ezban 《Biochemistry》1991,30(6):1533-1537
By the use of recombinant technology, several stable Chinese hamster ovary (CHO) cell lines expressing human FVIII were established. Thrombin treatment and SDS-PAGE analysis of the purified recombinant FVIII (rFVIII) revealed a striking difference from plasma-derived FVIII (pFVIII). A 43-kDa fragment of the FVIII heavy chain appears as a double band from rFVIII, while a single band from pFVIII is observed. All other fragments from the two samples appeared similar by SDS-PAGE. The heterogeneity is caused by incomplete tyrosine sulfation of one or more of the three potential tyrosine sulfation sites (Tyr718, Tyr719, Tyr723). To investigate if there is a general limitation and heterogeneity in the tyrosine sulfation of rFVIII, two other potential tyrosine sulfation sites on the FVIII light chain (Tyr1664, Tyr1680) were analyzed. The results show that both sites on the pFVIII light chain and on the rFVIII light chain are completely sulfated. The limitation of CHO cells to tyrosine sulfate rFVIII is therefore only restricted to a few sites. The two sulfated forms of rFVIII can easily be separated by ion-exchange chromatography, indicating the importance of the sulfate groups on the charge and/or conformation of FVIII. Both forms of rFVIII possess identical in vitro coagulation activity, von Willebrand factor binding, and thrombin activation profile. However, the difference in tyrosine sulfation may change other biological properties of FVIII.  相似文献   

15.
Radioimmunoassays for anionic and cationic dog trypsins are described. Characterization of the immunoreactivities in sera from fasting dogs demonstrated the presence of the two proenzymes only. Fasting sera from 10 dogs contained anionic and cationic trypsinogen in concentrations between 17-110 micrograms/l and 7-19 micrograms/l, respectively. Induction of experimental pancreatitis in dogs was accompanied by a large increase of immunoreactive anionic and cationic trypsins in the circulation. During the progress of the pancreatitis, immunoreactive trypsin with larger molecular weight than trypsinogen appeared. This high molecular weight immunoreactive trypsin was not seen in serum after intravenous injection of pancreatic juice in dogs. The high molecular weight immunoreactive trypsin probably consists of trypsin in complex with protease inhibitors. In vitro studies showed that the immunoreactivity of trypsin decreased considerably after binding to alpha 1-antitrypsin or alpha-macroglobulins.  相似文献   

16.
The alpha-chain of the fourth component of complement (C4) contains tyrosine sulfate (Karp, D.R. (1983) J. Biol. Chem. 258, 12745-12748). Here we have determined the site and stoichiometry of sulfation of C4 secreted by the human hepatoma-derived cell line Hep G2. C4 was labeled with [35S]sulfate and isolated from culture medium by immunoprecipitation. C4 digested with trypsin and chymotrypsin and analyzed by reverse-phase high-performance liquid chromatography contained a single sulfate-labeled peptide. Digestion of C4 with trypsin alone yielded two major sulfate-labeled peptides, suggesting that there may be some sequence variability in C4 near the site of sulfation. Sequential Edman degradation of tryptic peptides labeled with [3H]tyrosine and [35S]sulfate detected tyrosine residues at positions 5, 13, 16, and 18. Chymotrypsin cleaved 5 residues off the NH2-terminal end of tryptic peptides, yielding a peptide with tyrosine at positions 8, 11, and 13. Comparison of the position of tyrosine residues with the reported sequence of C4 identified the sites of sulfation as tyrosine residues at positions 738, 741, and 743 in the alpha-chain of C4. All 3 of these tyrosine residues appeared to be sulfated. When sulfation of C4 was partially inhibited by addition of catechol to culture medium, three different forms of the peptide were resolved by high-performance liquid chromatography, consistent with peptides containing 1, 2, or 3 sulfates. Comparison of the quantities of tyrosine and tyrosine sulfate in C4 which had been labeled with [3H]tyrosine and digested with Pronase also indicated that C4 contained an average of 2-3 residues of tyrosine sulfate/molecule. These results suggest that the biologically active form of the protein is sulfated.  相似文献   

17.
18.
In the absence of changes in functional mRNA levels, stimulation of the pancreas with caerulein, a peptide analog of cholecystokinin, has been previously shown to increase the synthesis of anionic but not cationic trypsinogen. To look for structure-function correlations, a high-yield, full-length cDNA library has been constructed from canine pancreatic poly(A)+ mRNA. Full-length clones coding for the two major trypsinogen isoenzyme forms have been identified by colony hybridization and verified by in vitro translation of hybrid-selected mRNA in the presence of microsomal membranes and an optimal redox potential. Disulfide-bonded translation products were separated and identified by two-dimensional isoelectric focusing-sodium dodecyl sulfate-gel electrophoresis. Nucleotide sequence analysis allowed us to deduce the amino acid sequences for the anionic and cationic forms of canine trypsinogen, which contain 232 and 231 residues, respectively (77% amino acid identity), and the 15-residue amino terminal signal sequences (53% amino acid identity) associated with the two presecretory forms. Measurements of relative and absolute mRNA levels, when related to relative protein synthesis values, indicated that the translational efficiency of anionic trypsinogen mRNA exceeded that of cationic trypsinogen mRNA by 1.5- to 2.9-fold under basal conditions. Analysis of the 5' noncoding regions of trypsinogen mRNAs revealed a striking conservation of sequence (10 of 12 bases) between dog and rat anionic trypsinogen forms. This contrasted markedly with the divergence of the 5' noncoding regions observed between dog anionic and cationic trypsinogen mRNAs.  相似文献   

19.
Mutations in the human cationic trypsinogen are associated with hereditary pancreatitis. The cDNA coding for human cationic trypsinogen was subcloned into the expression vector pcDNA3. The mutations R122H, N29I, A16V, D22G, and K23R were introduced by site directed mutagenesis. We constructed an expression vector coding for active trypsin by subcloning the cDNA of trypsin lacking the coding region for the trypsin activating peptide behind an appropriate signal peptide. Expression of protein was verified by Western blot and measurement of enzymatic activity. AR4-2J cells were transiently transfected with the different expression vectors and cell viability and intracellular caspase-3 activity were quantified. In contrast to wild-type trypsinogen, expression of active trypsin and mutated trypsinogens reduced cell viability of AR4-2J cells. Expression of trypsin and R122H trypsinogen induced caspase-3 activity. Acinar cells might react to intracellular trypsin activity by triggering apoptosis.  相似文献   

20.
Chymotrypsin C (CTRC) is a proteolytic regulator of trypsinogen autoactivation in humans. CTRC cleavage of the trypsinogen activation peptide stimulates autoactivation, whereas cleavage of the calcium binding loop promotes trypsinogen degradation. Trypsinogen mutations that alter these regulatory cleavages lead to increased intrapancreatic trypsinogen activation and cause hereditary pancreatitis. The aim of this study was to characterize the regulation of autoactivation of mouse trypsinogens by mouse Ctrc. We found that the mouse pancreas expresses four trypsinogen isoforms to high levels, T7, T8, T9, and T20. Only the T7 activation peptide was cleaved by mouse Ctrc, causing negligible stimulation of autoactivation. Surprisingly, mouse Ctrc poorly cleaved the calcium binding loop in all mouse trypsinogens. In contrast, mouse Ctrc readily cleaved the Phe-150–Gly-151 peptide bond in the autolysis loop of T8 and T9 and inhibited autoactivation. Mouse chymotrypsin B also cleaved the same peptide bond but was 7-fold slower. T7 was less sensitive to chymotryptic regulation, which involved slow cleavage of the Leu-149–Ser-150 peptide bond in the autolysis loop. Modeling indicated steric proximity of the autolysis loop and the activation peptide in trypsinogen, suggesting the cleaved autolysis loop may directly interfere with activation. We conclude that autoactivation of mouse trypsinogens is under the control of mouse Ctrc with some notable differences from the human situation. Thus, cleavage of the trypsinogen activation peptide or the calcium binding loop by Ctrc is unimportant. Instead, inhibition of autoactivation via cleavage of the autolysis loop is the dominant mechanism that can mitigate intrapancreatic trypsinogen activation.  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