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 328 毫秒
1.
Unlike bovine cationic trypsin, rat anionic trypsin retains activity at high pH. This alkaline stability has been attributed to stabilization of the salt bridge between the N-terminal Ile16 and Asp194 by the surface negative charge (Soman K, Yang A-S, Honig B, Fletterick R., 1989, Biochemistry 28:9918-9926). The formation of this salt bridge controls the conformation of the activation domain in trypsin. In this work we probe the structure of rat trypsinogen to determine the effects of the surface negative charge on the activation domain in the absence of the Ile16-Asp194 salt bridge. We determined the crystal structures of the rat trypsin-BPTI complex and the rat trypsinogen-BPTI complex at 1.8 and 2.2 A, respectively. The BPTI complex of rat trypsinogen resembles that of rat trypsin. Surprisingly, the side chain of Ile16 is found in a similar position in both the rat trypsin and trypsinogen complexes, although it is not the N-terminal residue and cannot form the salt bridge in trypsinogen. The resulting position of the activation peptide alters the conformation of the adjacent autolysis loop (residues 142-153). While bovine trypsinogen and trypsin have similar CD spectra, the CD spectrum of rat trypsinogen has only 60% of the intensity of rat trypsin. This lower intensity most likely results from increased flexibility around two conserved tryptophans, which are adjacent to the activation domain. The NMR spectrum of rat trypsinogen contains high field methyl signals as observed in bovine trypsinogen. It is concluded that the activation domain of rat trypsinogen is more flexible than that of bovine trypsinogen, but does not extend further into the protein core.  相似文献   

2.
The three-dimensional structure of the proteic complex formed by bovine trypsinogen and the porcine pancreatic secretory trypsin inhibitor (Kazal type) has been solved by means of Patterson search techniques, using a predicted model of the trypsin-ovomucoid complex (Papamokos et al., 1982). The structure of the complex, including 162 solvent molecules, has been refined at 1.8 Å resolution (26,341 unique reflections) to a conventional crystallographic R factor of 0.195. The inhibitor molecule binds to trypsinogen via hydrogen bonds and/or apolar interactions at sites P9, P7, P6, P5, P3, P1, P1′, P2′ and P3′ of the contact area. The structure of the inhibitor itself resembles closely that of the third domain of Japanese quail ovomucoid inhibitor, recently reported by Weber et al. (1981). The trypsinogen part of the complex resembles trypsin, as is the case in the trypsinogen-basic pancreatic trypsin inhibitor complex, but two segments of the activation domain adopt a different conformation. Most notably in the N-terminal region the Ile16-Gly19 loop, which is disordered in free trypsinogen and in the trypsinogen-basic pancreatic trypsin inhibitor complex (Huber & Bode, 1978), assumes a regular structure and the polypeptide chain can be traced as far as residue Asp14. This new and fixed structure allows the formation of a buried salt link between the side-chains of Lys156 and Asp194. Conformations differing from those of trypsin are also found for residues 20 to 28 and residues 141 to 155. Some structural perturbation is observed in other parts of the molecule, including the calcium loop.  相似文献   

3.
X-ray intensity data to 1.8 Å resolution were collected from native trigonal crystals of bovine trypsinogen. The orientation and position of the trypsinogen molecules within their crystal cells were determined by Patterson search techniques using the refined model of bovine trypsin (Bode &; Schwager, 1975), and by subsequent R factor refinement. The translation functions allowed discrimination between the enantiomorphic space groups P3221 and P3121. After one constrained crystallographic refinement cycle, which reduced the crystallographic reliability factor (R) from 35% to 31%, a preliminary difference Fourier map showed several interesting details. Several refinement cycles reduced the value of R to 23%. The overall chain folding is very similar to trypsin. The chain segments, including residues 184 to 1932 and 217 to 223, which form the specificity pocket in trypsin, are flexible in trypsinogen. The autolysis loop is partially mobile between residues 142 and 152. There is no continuing electron density for the N terminal residues preceding Tyr20. This indicates that the N terminus may be only weakly fixed to the rest of the molecule or may even float freely in solution.  相似文献   

4.
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   

5.
Trypsin-like serine proteases play essential roles in diverse physiological processes such as hemostasis, apoptosis, signal transduction, reproduction, immune response, matrix remodeling, development, and differentiation. All of these proteases share an intriguing activation mechanism that involves the transition of an unfolded domain (activation domain) of the zymogen to a folded one in the active enzyme. During this conformational change, activation domain segments move around highly conserved glycine hinges. In the present study, hinge glycines were replaced by alanine residues via site directed mutagenesis. The effects of these mutations on the interconversion of the zymogen-like and active conformations as well as on catalytic activity were studied. Mutant trypsins showed zymogen-like structures to varying extents characterized by increased flexibility of some activation domain segments, a more accessible N-terminus and a deformed substrate binding site. Our results suggest that the trypsinogen to trypsin transition is hindered by the mutations, which results in a shift of the equilibrium between the inactive zymogen-like and active enzyme conformations toward the inactive state. Our data also showed, however, that the inactive conformations of the various mutants differ from each other. Binding of substrate analogues shifted the conformational equilibrium toward the active enzyme since inhibited forms of the trypsin mutants showed similar structural features as the wild-type enzyme. The catalytic activity of the mutants correlated with the proper conformation of the active site, which could be supported by varying conformations of the N-terminus and the autolysis loop. Transient kinetic measurements confirmed the existence of an inactive to active conformational transition occurring prior to substrate binding.  相似文献   

6.
Two monoclonal antibodies (Mab) raised against human pancreatic trypsin 1, Mab G6 and A8, were previously isolated and characterized. The two Mab which recognize trypsinogen 1 are found to inhibit the activation of trypsinogen 1 by enterokinase. The inhibition of activation by the two Mab is concentration-dependent, rapid and virtually complete with Mab G6. Activation of trypsinogen 2 is totally inhibited by Mab G6, while Mab A8 has no effect on the activation of trypsinogen 2. The two monoclonal antibodies have opposite effects on the proteolytic activity of trypsin 1; Mab G6 increases proteolytic activity while Mab A8 inhibits trypsin activity by as much as 40%. This inhibition is concentration dependent but cannot account for the complete inhibition of activation of trypsinogen 1. Neither monoclonal antibody significantly inhibits the esterolytic activity of either form of human trypsin. Western-blot analysis of the reactivity of the two monoclonal antibodies with trypsinogens of various species shows that only Mab G6 cross-reacts with dog trypsinogen.  相似文献   

7.
The complex formed by bovine trypsinogen and the pancreatic trypsin inhibitor crystallizes in large crystals isomorphous with trypsin-PTI2 complex crystals Rühlmann et al. 1973. X-ray diffraction data to 1.9 Å resolution were collected in the absence and presence of Ile-Val dipeptide. Both trypsinogen complex structures have been crystallographically refined, using the refined trypsin-PTI complex Huber et al. 1974a as a starting model. The final R values are 0.25 and 0.26, respectively. The mean main-chain atom deviations between the three complex structures are about 0.15 Å. In contrast, the mean deviation between the complexed and the free trypsinogen Fehlhammer et al. 1977 is 0.28 Å, reflecting the influence of crystal packing and complexation. The trypsinogen component adopts a trypsin-like conformation upon PTI binding: The Asp194 side-chain turns around and the activation domain becomes rigid, forming the specificity pocket and the Ile16 binding cleft. The specific interactions between PTI and trypsin are also observed in the trypsinogen complex. As in free trypsinogen, the N-terminus including residues Val10 to Gly18 is mobile and sticks out into solution. Apart from the different arrangement of the N-termini in the two complexes, the only significant, but minor structural difference is the enhanced thermal mobility of the autolysis loop in the trypsinogen complex. Upon binding of the Ile-Val dipeptide, the autolysis loop becomes fixed as in the trypsin complex. The Ile-Val position is identical in the ternary and the trypsin complex.  相似文献   

8.
A trypsin was purified from the hepatopancreas of snakehead (Channa argus) by ammonium sulfate fractionation and a series of column chromatographies including DEAE-Sepharose, Sephacryl S-200 HR and Hi-Trap Capto-Q. The molecular mass of the purified trypsin was about 22 kDa, as estimated by SDS-PAGE. The optimum pH and temperature of the purified trypsin were 9.0 and 40 °C, respectively. The trypsin was stable in the pH range of 7.5-9.5 and below 45 °C. The enzymatic activity was strongly inhibited by serine proteinase inhibitors, such as MBTI, Pefabloc SC, PMSF, LBTI and benzamidine. Peptide mass fingerprinting (PMF) of the purified protein obtained 2 peptide fragments with 25 amino acid residues and were 100% identical to the trypsinogen from pufferfish (Takifugu rubripes). The activation energy (Ea) of this enzyme was 24.65 kJ·M− 1. Apparent Km was 1.02 μM and kcat was 148 S− 1 for fluorogenic substrate Boc-Phe-Ser-Arg-MCA. A trypsinogen gene encoding 247 amino acid residues was further cloned on the basis of the sequence obtained from PMF and the conserved site peptide of trypsinogen together with 5′-RACE and 3′-RACE. The deduced amino acid sequence contains a signal peptide of 15 residues and an activation peptide of 9 amino acid residues with a mature protein of 223 residues. The catalytic triad His-64, Asp-107, Ser-201 and 12 Cys residues which may form 6 disulfide bonds were conserved. Compared with the PMF data, only 2 amino acid residues difference were identified, suggesting the cloned trypsinogen is quite possibly the precursor of the purified trypsin.  相似文献   

9.
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   

10.
11.
The amino-terminal sequence of the catalytic subunit of bovine enterokinase   总被引:2,自引:0,他引:2  
Bovine enterokinase (enteropeptidase) is a serine protease and functions as the physiological activator of trypsinogen. The enzyme has a heavy chain (115 kD) covalently linked to a light or catalytic subunit (35 kD). The amino acid composition showed that the light chain has nine half-cystine residues (four as intramolecular disulfides) and that one half-cystine was in a disulfide link between the light and heavy subunits. The amino-terminal 27 residues of the S-vinylpyridyl derivative of the light chain were determined by gas-phase Edman degradation. The sequence has homologies with other serine proteases containing one or two chains. The homologies suggest that the catalytic subunit has the same three-dimensional structure and, therefore, the same mechanism of enzymatic action as pancreatic chymotrypsin, trypsin, and elastase. The presence of the conserved amino-terminal activation peptide sequence (IVGG) shows that enterokinase must have a zymogen precursor and that the two-chain enzyme arises from limited proteolysis during posttranslational processing.  相似文献   

12.
The activation of trypsinogen to trypsin in the small intestine can occur by the action of enterokinase or, alternatively, as an autocatalytic process catalysed by trypsin itself. We have found that bile salts and human bile cause a significant enhancement of the autocatalytic activation of trypsinogen. This effect is dependent on the calcium ion concentration and is most marked around pH 5.4 and 7.8. An optimum concentration exists for each bile salt at which the greatest enhancement occurs. At this concentration, certain bile salts have been shown to produce activation effects of up to 55-fold. It is suggested that this activation of the autocatalytic process by bile plays an important role in protein digestion in the small intestine, since it has been shown previously that duodenal trypsin levels are abnormally low in patients with an impairment of bile secretion.  相似文献   

13.
Mutations in human cationic trypsinogen (PRSS1) cause autosomal dominant hereditary pancreatitis. Increased intrapancreatic autoactivation of trypsinogen mutants has been hypothesized to initiate the disease. Autoactivation of cationic trypsinogen is proteolytically regulated by chymotrypsin C (CTRC), which mitigates the development of trypsin activity by promoting degradation of both trypsinogen and trypsin. Paradoxically, CTRC also increases the rate of autoactivation by processing the trypsinogen activation peptide to a shorter form. The aim of this study was to investigate the effect of CTRC on the autoactivation of clinically relevant trypsinogen mutants. We found that in the presence of CTRC, trypsinogen mutants associated with classic hereditary pancreatitis (N29I, N29T, V39A, R122C, and R122H) autoactivated at increased rates and reached markedly higher active trypsin levels compared with wild-type cationic trypsinogen. The A16V mutant, known for its variable disease penetrance, exhibited a smaller increase in autoactivation. The mechanistic basis of increased activation was mutation-specific and involved resistance to degradation (N29I, N29T, V39A, R122C, and R122H) and/or increased N-terminal processing by CTRC (A16V and N29I). These observations indicate that hereditary pancreatitis is caused by CTRC-dependent dysregulation of cationic trypsinogen autoactivation, which results in elevated trypsin levels in the pancreas.  相似文献   

14.
15.
16.
Intra-acinar cell nuclear factor-kappaB (NF-kappaB) and trypsinogen activation are early events in secretagogue-induced acute pancreatitis. We have studied the relationship between NF-kappaB and trypsinogen activation in rat pancreas. CCK analogue caerulein induces early (within 15 min) parallel activation of both NF-kappaB and trypsinogen in pancreas in vivo as well as in pancreatic acini in vitro. However, NF-kappaB activation can be induced without trypsinogen activation by lipopolysaccharide in pancreas in vivo and by phorbol ester in pancreatic acini in vitro. Stimulation of acini with caerulein after 6 h of culture results in NF-kappaB but not trypsinogen activation. Protease inhibitors (AEBSF, TLCK, and E64d) inhibit both intracellular trypsin activity and NF-kappaB activation in caerulein stimulated acini. A chymotrypsin inhibitor (TPCK) inhibits NF-kappaB activation but not trypsin activity. The proteasome inhibitor MG-132 prevents caerulein-induced NF-kappaB activation but does not prevent trypsinogen activation. These findings indicate that although caerulein-induced NF-kappaB and trypsinogen activation are temporally closely related, they are independent events in pancreatic acinar cells. NF-kappaB activation per se is not required for the development of early acinar cell injury by supramaximal secretagogue stimulation.  相似文献   

17.
Several enzymic and physical properties of Sepharose-bound trypsin and activated Sepharose-bound trypsinogen have been compared to those of the soluble enzyme. Sepharose-bound trypsinogen could be activated to the same extent as soluble trypsinogen; the release of the activation peptide and formation of the active site occurred as expected in the presence of catalytic amounts of trypsin. With synthetic substrates, the relative activity and pH dependence of both immobilized trypsin preparations were essentially identical and nearly the same as the soluble enzyme. Sepharose-trypsin also formed an inactive complex with soybean trypsin inhibitor, with 85% of the active sites participating. In contrast, the activity of Sepharose-trypsin with chymotrypsinogen and with trypsinogen as substrates was only 40% that of soluble trypsin. There is evidence for some catalytic heterogeneity of active sites of bound trypsin; probably those sites buried within the gel have a limited catalytic efficiency with macromolecular substrates. The immobilized enzyme is more stable than the soluble enzyme at elevated temperatures and to concentrated urea, and denaturation by urea at pH 8 is fully reversible since the loss of molecules by autolysis is eliminated.  相似文献   

18.
1. A powerful kinase which changes trypsinogen to trypsin was found to be present in the synthetic liquid culture medium of a mold of the genus Penicillium. 2. The concentration of kinase in the medium is increased gradually during the growth of the mold organism and continues to increase for some time even after the mold has ceased growing. 3. Mold kinase transforms trypsinogen to trypsin only in an acid medium. It differs thus from enterokinase and trypsin which activate trypsinogen best in a slightly alkaline medium. 4. The action of the mold kinase in the process of transformation of trypsinogen is that of a typical enzyme. The process follows the course of a catalytic unimolecular reaction, the rate of formation of a definite amount of trypsin being proportional to the concentration of kinase added. The ultimate amount of trypsin formed, however, is independent of the concentration of kinase used. 5. The formation of trypsin from trypsinogen by mold kinase is not accompanied by any measurable loss of protein. 6. The temperature coefficient of formation of trypsin from trypsinogen by mold kinase varies from Q 5–15 = 1.70 to Q 25–30 = 1.25 with a corresponding variation in the value of µ from 8100 to 4250. 7. Trypsin formed from trypsinogen by means of mold kinase is identical in crystalline form with the crystalline trypsin obtained by spontaneous autocatalytic activation of trypsinogen at pH 8.0. The two products have within the experimental error the same solubility and specific activity. A solution saturated with the crystals of either one of the trypsin preparations does not show any increase in protein concentration or activity when crystals of the other trypsin preparation are added. 8. The Penicillium mold kinase has a slight activating effect on chymo-trypsinogen the rate being only 1–2 per cent of that of trypsinogen. The activation, as in the case of trypsinogen, takes place only in an acid medium. 9. Mold kinase is rapidly destroyed when brought to pH 6.5 or higher, and also when heated to 70°C. In the temperature range of 50–60°C. the inactivation of kinase follows a unimolecular course with a temperature coefficient of Q 10 = 12.1 and µ = 53,500. The molecular weight of mold kinase, as determined by diffusion, is 40,000.  相似文献   

19.
Proteins with trypsin-like immunoreactivity (first detected by a specific immunoenzymatic assay) were isolated from CAPAN-1 and CFPAC-1 cell culture-conditioned media by chromatography on an immunoadsorbent prepared with a polyclonal antibody directed against trypsin 1. The adsorbed proteins were devoid of free trypsin activity but trypsin activity was present after enterokinase activation demonstrating that the immunoreactive trypsin present in cell supernatants corresponds to trypsinogens. When characterised by Western blotting using a monoclonal antibody directed against human trypsin 1 two protein bands corresponding to trypsinogen 1 (23 kDa) and trypsinogen 2 (25 kDa) gave a positive reaction. These results demonstrate the presence of trypsinogens 1 and 2 in CAPAN-1 and CFPAC-1 cells and in their culture-conditioned media.  相似文献   

20.
One of ostrich (Struthio camelus) trypsinogen genes was cloned from pancreatic cDNA. Its amino acid sequence compared to known trypsin sequences from other species shows high identity and suggests that it is a member of the phylogenetically anionic trypsinogen I subfamily. After cytoplasmic over expression in Escherichia coli and renaturation, the activation properties of ostrich trypsinogen were studied and compared to those of human trypsinogen 1 (also called as human cationic trypsinogen). Ostrich trypsinogen undergoes bovine enterokinase activation and autoactivation much faster than human trypsinogen 1 and exhibits on a synthetic substrate a somewhat higher enzymatic activity than the latter one. The most interesting property of ostrich trypsin is its relatively fast autolysis that can be explained via a mechanism different from the common mechanism for rat and human 1 trypsins. The latter proteases have a site, Arg117-Val118, where the autolysis starts and then goes on in a zipper-like fashion. This is absent from ostrich trypsin. Instead it has a couple of cleavage sites within regions 67-98, including two unusual ones, Arg76-Glu77 and Arg83-Ser84. These appear to be hydrolysed fast in a non-consecutive manner. Such an autolysis mechanism could not be inhibited by a single-site mutation which in humans is proposed to lead to pancreatitis.  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