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.
Pepsin is formed as the zymogen, pepsinogen, which includes an additional 44 residue prosegment (PS) on the N-terminus. Upon acidification (pH <3) the PS is removed, yielding active pepsin. The PS is critical to such processes as the initiation of correct folding and protein stability. In the present study, the NMR assignments of the 34.6 kDa native porcine pepsin and porcine pepsin complexed with pepstatin are reported in order to obtain structural information regarding PS-catalyzed protein folding. Such information would contribute to a better understanding of the nature of folding/unfolding energy barrier of pepsin and other aspartic proteases.  相似文献   

2.
This paper presents a new system for the soluble expression and characterization of porcine pepsinogen from the methylotrophic yeast Pichia pastoris. The cDNA that encodes the zymogenic form of porcine pepsin (EC 3.4.23.1) was cloned into the EcoRI site of the vector pHIL-S1 downstream from the AOX1 alcohol oxidase promoter. After P. pastoris transformation, colonies were screened for expression of pepsinogen based on enzyme activity of the active form, pepsin. The recombinant enzyme was purified 138-fold by anion exchange and affinity column chromatography. Homogeneity was confirmed through SDS-PAGE, Western blot, and N-terminal sequencing. When compared to commercial pepsin, the recombinant pepsin had similar kinetic profiles, pH/temperature stability, and secondary/tertiary conformation. A glycosylated form was also isolated and found to exhibit kinetic and structural characteristics similar to those of the commercial and wild-type pepsin, but was slightly more thermal stable. The above results indicate that the P. pastoris expression system offers a convenient and efficient means to produce and purify a soluble form of pepsin(ogen).  相似文献   

3.
In order to increase understanding of the basis of the stability of the native conformational state of porcine pepsin A, a strategy based on induction and monitoring of protein denaturation was developed. Structural perturbation was achieved by adding acetonitrile (MeCN) to the protein-solvent system. MeCN was found to induce non-coincident disruption of the secondary and tertiary structural features of pepsin A. It is proposed that gross unfolding is prompted by disruption of the protein hydration pattern induced by the organic co-solvent. It should be noted that the functional properties and thermal stability of the protein were already impaired before the onset of global unfolding. Low and intermediate contents of MeCN in the protein-solvent system affected the sharpness of the thermal transition and the degree of residual structure of the heat-denatured state. The importance of hydration to the conformational stability of pepsin A in its biologically active state is discussed.  相似文献   

4.
The molecular structure of the archetypal aspartic proteinase, porcine pepsin (EC 3.4.23.1), has been refined using data collected from a single monoclinic crystal on a twin multiwire detector system to 1.8 A resolution. The current crystallographic R-factor (= sigma parallel to Fo/-/Fc parallel to/sigma/Fo/) is 0.174 for the 20,519 reflections with /Fo/ greater than or equal to 3 sigma (Fo) in the range 8.0 to 1.8 A (/Fo/ and /Fc/ are the observed and calculated structure factor amplitudes respectively). The refinement has shown conclusively that there are only 326 amino acid residues in porcine pepsin. Ile230 is not present in the molecule. The two catalytic residues Asp32 and Asp215 have dispositions in porcine pepsin very similar to the dispositions of the equivalent residues in the other aspartic proteinases of known structure. A bound solvent molecule is associated with both carboxyl groups at the active site. No bound ethanol molecule could be identified conclusively in the structure. The average thermal motion parameter of the residues that comprise the C-terminal domain of pepsin is approximately twice that of the residues in the N-terminal domain. Comparisons of the tertiary structure of pepsin with porcine pepsinogen, penicillopepsin, rhizopus pepsin and endothia pepsin reveal that the N-terminal domains are topographically more similar than the conformationally flexible C-terminal domains. The conformational differences may be modeled as rigid-body movements of "reduced" C-terminal domains (residues 193 to 212 and 223 to 298 in pepsin numbering). A similar movement of the C-terminal domain of endothia pepsin has been observed upon inhibitor binding. A phosphoryl group covalently attached to Ser68 O gamma has been identified in the electron density map of porcine pepsin. The low pKa1 value for this group, coupled with unusual microenvironments for several of the aspartyl carboxylate groups, ensures a net negative charge on porcine pepsin in a strongly acid medium. Thus, there is a structural explanation for the very early observations of "anodic migration" of porcine pepsin at pH 1. In the crystals, the molecules are packed tightly into a monoclinic unit cell. There are 190 direct contacts (less than or equal to 4.0 A) between a central pepsin molecule and the five unique symmetry-related molecules surrounding it in the crystalline lattice. The tight packing in this cell makes pepsin's active site and binding cleft relatively inaccessible to substrate analogs or inhibitors.  相似文献   

5.
It was the aim of this study to evaluate the stability of two novel cystine-knot microproteins (CKM) SE-ET-TP-020 and SE-MC-TR-020 with potential clinical relevance towards luminally secreted proteases of the gastrointestinal tract in order to gain information about their potential for oral administration. Therefore, the stability of the two CKM and the model-drug insulin towards collected porcine gastric and small intestinal juice as well as towards isolated proteolytic enzymes was evaluated under physiological conditions. No intact SE-ET-EP-020 was detected after few seconds of incubation with porcine small intestinal juice. SE-ET-TP-020 was also degraded in porcine gastric juice. Furthermore, SE-ET-TP-020 was extensively degraded by isolated chymotrypsin, trypsin and pepsin. Moreover, it was degraded by elastase. SE-MC-TR-020 was degraded entirely within approximately 2 h when incubated in porcine small intestinal juice, whereas no degradation was observed within a 3 h incubation period with porcine gastric juice. In presence of the isolated proteolytic enzymes, SE-MC-TR-020 was only slightly degraded by trypsin and pepsin, whereas elastase caused no degradation to SE-MC-TR-020 at all. Chymotrypsin was the protease that caused most degradation to SE-MC-TR-020. The model drug insulin was degraded extensively by chymotrypsin, elastase, pepsin and trypsin as well as by porcine gastric and porcine small intestinal juice. In conclusion, a precise characterisation of SE-ET-TP-020 and SE-MC-TR-020 degrading luminally secreted GI enzymes has been made, which is an important and substantial prerequisite for the further optimisation of these CKM.  相似文献   

6.
A study was undertaken to examine the effects of N-linked glycosylation on the structure-function of porcine pepsin. The N-linked motif was incorporated into four sites (two on the N-terminal domain and two on the C-terminal domain), and the recombinant protein expressed using Pichia pastoris. All four N-linked recombinants exhibited similar secondary and tertiary structure to nonglycosylated pepsin, that is, wild type. Similar K(m) values were observed, but catalytic efficiencies were approximately one-third for all mutants compared with the wild type; however, substrate specificity was not altered. Activation of pepsinogen to pepsin occurred between pH 1.0 to 4.0 for wild-type pepsin, whereas the glycosylated recombinants activated over a wider range, pH 1.0 to 6.0. Glycosylation on the C-terminal domain exhibited similar pH activity profiles to nonglycosylated pepsin, and glycosylation on the N-domain resulted in a change in activity profile. Overall, glycosylation on the C-domain led to a more global stabilization of the structure, which translated into enzymatic stability, whereas on the N-domain, an increase in structural stability had little effect on enzymatic stability. Finally, glycosylation on the flexible loop region also appeared to increase the overall structural stability of the protein compared with wild type. It is postulated that the presence of the carbohydrate residues added rigidity to the protein structure by reducing conformational mobility of the protein, thereby increasing the structural stability of the protein.  相似文献   

7.
Pepsinogen was isolated from the gastric mucosa of Trimeresurus flavoviridis (Habu snake) by DEAE-cellulose and DEAE-Sepharose ion-exchange chromatographies, and Sephacryl S-200 gel-chromatography. The yield calculated from the crude extract was 29% with 6.2-fold purification. The purified pepsinogen gave a single band on both native- and SDS-PAGE. As no other active enzyme was detected on the chromatographies, it was concluded that the Habu snake has one major pepsinogen. The molecular mass of the pepsinogen was estimated to be 38 kDa by SDS-PAGE. The sequence of the N-terminal 26 amino acid residues was determined and compared with those of other pepsinogens. The N-terminal structure of Habu snake pepsinogen was more homologous with those of mammalian pepsinogens C than those of mammalian pepsinogens A. The pepsinogen was rapidly converted to pepsin by way of an intermediate form induced by acidification. The optimum pH of Habu snake pepsin for bovine hemoglobin was 1.5-2.0, and it retained full activity at pH 6.2 and 30 degrees C on incubation for 30 min. The optimum temperature for the snake pepsin was 50 degrees C and it was stable at 40 degrees C on incubation for 10 min. The proteolytic activity of the pepsin toward bovine hemoglobin was about two times higher than that of porcine pepsin A, however, the activity toward oxidized bovine insulin B-chain was lower than that of porcine pepsin A, and it did not hydrolyze oligopeptides. The specificity for oxidized bovine insulin B-chain of the pepsin was different from that of porcine pepsin A. Habu snake pepsin was inhibited by pepstatin A but not by serine, cysteine, or metallo protease inhibitors.  相似文献   

8.
1. The reactivities of pepsins, isolated from three different sources (porcine, bovine, and Penicillium jathiuelum), toward ester and peptide substrates were compared. 2. Porcine pepsin showed the highest activity followed by penicillopepsin with bovine pepsin being the least active. 3. The esterase activity of penicillopepsin was greater than that of porcine pepsin with bovine pepsin again showing the least activity. 4. The CD spectra indicate that porcine and bovine pepsin have similar conformations, even though bovine pepsin shows less ellipticity at 220 nm. 5. Penicillopepsin showed a completely opposite sign in the near-u.v. region of the CD spectrum. 6. The far-u.v. region of the CD spectrum of penicillopepsin strongly suggests a beta-sheet structure. 7. Previously reported X-ray crystallographic data suggest that porcine pepsin has a compact three-dimensional structure, while the structures of bovine and penicillopepsins are partially unfolded.  相似文献   

9.
The specific inhibitor, N-diazoacetylnorleucine methyl ester reacts stoicheiometrically with bovine pepsin resulting in a simultaneous loss of all enzymic activity. A peptide containing a modified aspartyl group was isolated from bovine pepsin labelled with (14)C-labelled inhibitor. The aspartic acid residue is presumed to be part of the active centre and is in the same heptapeptide sequence as in porcine pepsin: Ile-Val-Asp-Thr-Gly-Thr-Ser.  相似文献   

10.
A limited proteolysis of bovine pepsin (EC 3.4.4.1) was carried out. A proteolysis-resistant C-terminal protein fragment containing about 170 amino acid residues was isolated and its N-terminal sequence was established, using Edman's automatic method. It was assumed that the fragment of bovine pepsin isolated, similar to the previosly obtained porcine pepsin fragment, is an independent constituent of the protein molecule.  相似文献   

11.
A standard immunofluorescent method was modified for the staining of leptospires in formalin fixed, paraffin embedded tissues. Routine histologic sections were deparaffinized and treated with pepsin prior to staining. Pepsin treatment greatly enhanced subsequent staining of leptospires in naturally infected bovine and porcine tissues as well as in artificially infected tissues. Leptospires in naturally infected bovine tissues were usually undetectable in untreated sections but clearly visible in stained pepsin-treated sections. Naturally infected porcine kidney usually contained high levels of leptospiral antigen which could be stained without prior pepsin treatment. However, pepsin treatment of porcine tissues greatly increased the amount of leptospiral antigen detectable and made individual leptospires more conspicuous. The staining method could employ a single antiserum for the staining of leptospires from 13 serogroups. Also, leptospires could be stained in tissues stored in formalin for more than 14 months and in 26-year-old paraffin embedded tissues.  相似文献   

12.
A standard immunofluorescent method was modified for the staining of leptospires in formalin fixed, paraffin embedded tissues. Routine histologic sections were deparaffinized and treated with pepsin prior to staining. Pepsin treatment greatly enhanced subsequent staining of leptospires in naturally infected bovine and porcine tissues as well as in artificially infected tissues. Leptospires in naturally infected bovine tissues were usually undetectable in untreated sections but clearly visible in stained pepsin-treated sections. Naturally infected porcine kidney usually contained high levels of leptospiral antigen which could be stained without prior pepsin treatment. However, pepsin treatment of porcine tissues greatly increased the amount of leptospiral antigen detectable and made individual leptospires more conspicuous. The staining method could employ a single antiserum for the staining of leptospires from 13 serogroups. Also, leptospires could be stained in tissues stored in formalin for more than 14 months and in 26-year-old paraffin embedded tissues.  相似文献   

13.
A crude extract of the proventriculus of the Japanese quail gave at least five bands of peptic activity at pH 2.2 on polyacrylamide gel electrophoresis. The main component, constituting about 40% of the total acid protease activity, was purified to homogeneity by hydroxyapatite and DEAE-Sepharose column chromatographies. At below pH 4.0, the pepsinogen was converted to a pepsin, which had the same electrophoretic mobility as one of the five bands of peptic activity present in the crude extract. The molecular weights of the pepsinogen and the pepsin were 40 000 and 36 000, respectively. Quail pepsin was stable in alkali up to pH 8.5. The optimal pH of the pepsin on hemoglobin was pH 3.0. The pepsin had about half the milk-clotting activity of purified porcine pepsin, but the pepsinogen itself had no activity. The hydrolytic activity of quail pepsin on N-acetyl-L-phenylalanyl-3,5-diiodo-L-tyrosine was about 1% of that of porcine pepsin. Among the various protease inhibitors tested, only pepstatin inhibited the proteolytic activity of the pepsin. The amino acid composition of quail pepsinogen was found to be rather similar to that of chick pepsinogen C, and these two pepsinogens possessed common antigenicity.  相似文献   

14.
The activation of Sepharose-bound monkey pepsinogen A under acidic conditions proceeded by cleavage of the Leu47-Ile48 bond, indicating the occurrence of the intramolecular one-step activation, although the rate of cleavage was very slow. On the other hand the activation of monkey pepsinogen A in solution was highly dependent on pepsinogen concentration and the addition of exogenous pepsin A accelerated the rate of activation, indicating the predominance of intermolecular reaction. The cleavage site, however, was also restricted to the Leu47-Ile48 bond. Thus, apparently exclusive one-step activation occurred in monkey pepsinogen. The activation of porcine pepsinogen A in solution was also dependent on pepsinogen concentration and the addition of exogenous pepsin A accelerated the rate of activation. The major cleavage site by the exogenously added pepsin was the Leu44-Ile45 bond. Therefore the site most susceptible to the intermolecular attacks was the bond connecting the activation segment and the pepsin moiety in both monkey and porcine pepsinogens. In porcine pepsinogen, however, a part of the zymogen was activated through the intermediate form, and an intramolecular reaction was suggested to be involved in the generation of this form. These results showed that in both pepsinogens A the intramolecular reaction occurred, first yielding pepsin A or the intermediate form, which then acted intermolecularly on the remaining pepsinogen or the intermediate form to complete the activation in a short time. A molecular mechanism for the activation reaction was proposed to explain consistently the experimental results.  相似文献   

15.
Limited proteolysis of bovine kappa-casein has been investigated with porcine pepsin A and C, and with the 2 microbial proteinases Mucor miehei proteinase and Endothia parasitica proteinase. The liberated C-terminal glycomacropeptide of kappa-casein was isolated after precipitation in 3% trichloroacetic acid followed by high-performance gel-permeation chromatography on a TSK G3000 SW column. From amino acid analyses and N-terminal sequencing of the liberated peptide it is concluded that porcine pepsin A, C and Mucor miehei proteinase cleave the same bond as chymosin: Phe-105-Met-106 whereas Endothia parasitica proteinase cleaves the bond Ser-104-Phe-105.  相似文献   

16.
Pepsin, a member of the aspartate protease family, exists in a partially unfolded state at alkaline pH where the N-terminal domain of pepsin has a flexible structure while the C-terminal domain has a highly folded structure. In this work, the conformational stability of porcine pepsin in an alkaline denatured (A(D)) state against acetonitrile and ethanol solvents was studied using a combination of electronic circular dichroism (ECD) and fluorescence techniques. The ECD results demonstrate that both ethanol and acetonitrile induce secondary structural changes in pepsin at A(D) state. However, the minimum concentration required to induce significant secondary structural changes in pepsin varies for ethanol (>30%, v/v) and acetonitrile (>60%, v/v) solvents. At maximum concentration used (90%, v/v), both solvents induce predominantly β-sheet conformation. Unlike acetonitrile, ethanol induces significant amount of non-native α-helical conformations at the intermediate concentrations (50-80%). The tryptophan fluorescence results demonstrate that both acetonitrile and ethanol induce substantial changes in the tertiary structure of pepsin in the A(D) state above certain concentrations. The current results have important implications in understanding the effect of co-solvents on the conformation of proteins in the "denatured state".  相似文献   

17.
Amino acid sequence of porcine spleen cathepsin D light chain   总被引:5,自引:0,他引:5  
The complete amino acid sequence of the light chain of cathepsin D from porcine spleen has been determined. The light chain consists of a single polypeptide chain with 97 amino acid residues. The sequence is: (formula; see text) The molecular weight of the light chain was calculated from this sequence to be 10,548 (without carbohydrates). A single disulfide bond links two half-cystine residues between positions 46 and 53. A cysteine residue is located at position 27. The light chain sequence is extensively homologous to the NH2-terminal sequence of other aspartyl proteases. It shows a 59% identity with the sequence of mouse submaxillary gland renin and a 49% identity with that of porcine pepsin. A single glycosylation site is located at residue 70 of the cathepsin D light chain. This site corresponds to position 67 of pepsin by homology. The active site aspartyl residue, corresponding to Asp-32 of pepsin, is located at residue 33 in the cathepsin D light chain.  相似文献   

18.
Dee D  Pencer J  Nieh MP  Krueger S  Katsaras J  Yada RY 《Biochemistry》2006,45(47):13982-13992
A zymogen-derived protein, pepsin, appears to be incapable of folding to the native state without the presence of the prosegment. To better understand the nature of the irreversible denaturation of pepsin, the present study reports on the characterization of the stability and low-resolution tertiary and secondary structures of native, alkaline unfolded and acid refolded porcine pepsin. Through a combination of small-angle neutron scattering (SANS), CD, and DSC, acid refolded pepsin (Rp) was shown to have secondary and tertiary structures intermediate between the alkaline denatured and native forms but was found to be thermodynamically stable relative to the native state. It was also observed that the acid refolded state of pepsin was dependent on the protein concentration during refolding because CD and SANS data revealed that both the secondary and tertiary structures of concentrated-refolded pepsin (>10 mg/mL) (CRp) were native-like, in contrast to the intermediate nature of Rp, refolded under dilute concentration (<10 mg/mL). Despite a native-like conformation, CRp was more stable and had substantially reduced activity compared to that of the native state, suggesting that the protein was misfolded. It is proposed that the stable but misfolded, acid-refolded states are evidence that pepsin in its native conformation was metastable. Furthermore, the disruption of the active site cleft in the denatured states could be discerned by modeling of the SANS data.  相似文献   

19.
In this study the ability of three polyamidoamine (PAMAM) dendrimers with different surface charge (positive, neutral and negative) to interact with a negatively charged protein (porcine pepsin) was examined. It was shown that the dendrimer with a positively charged surface (G4 PAMAM-NH2), as well as the dendrimer with a neutral surface (G4 PAMAM-OH), were able to inhibit enzymatic activity of pepsin. It was also found that these dendrimers act as mixed partially non-competitive pepsin inhibitors. The negatively charged dendrimer (G3.5 PAMAM-COOH) was not able to inhibit the enzymatic activity of pepsin, probably due to the electrostatic repulsion between this dendrimer and the protein. No correlation between changes in enzymatic activity of pepsin and alterations in CD spectrum of the protein was observed. It indicates that the interactions between dendrimers and porcine pepsin are complex, multidirectional and not dependent only on disturbances of the secondary structure.  相似文献   

20.
To investigate the active site structures of porcine pepsin and Rhizopus chinensis acid protease (RAP), spin label techniques were applied for these enzymes. Comparison of spin labeled porcine pepsin and RAP suggested that the active site cleft of porcine pepsin was narrower at the top, but wider at the bottom than that of RAP. Addition of pepstatin restricted the motion of the labeled nitroxide radicals. Under alkaline conditions, the enzymes changed their conformation discontinuously and irreversibly to open the active site clefts and to lose the binding ability for pepstatin. The denaturation points of both the enzymes were determined to be pH 6.2.  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