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.
Exposure of pepsinogen to acid for less than 2 min yields a product with proteolytic activity. This activity is due to intramolecular and intermolecular formation of pepsin from pepsinogen. We find no evidence for intermolecular proteolytic activity in the zymogen. These conclusions are based upon two sets of experiments. First, chemical cleavage of pepsinogen during short activation is demonstrated by quantitative analysis of the NH2-terminal 2 residues of the pepsin and pepsinogen in an activation mixture. In addition, quantitative NH2-terminal analyses after activation under different conditions confirm our previous inference that the product of unimolecular pepsinogen activation is homogeneous whereas bimolecular activation produces a pepsin product with a variety of NH2 termini. Second, spectral changes which occur upon acidification of a pepsinogen solution and are reversed by neutralization are shown to be consistent with the chemical cleavage of pepsinogen during acidification. The first order rate constant for pepsinogen activation, calculated from these spectral experiments, agrees well with the value we had determined previously.  相似文献   

2.
Most eukaryotic aspartic protease zymogens are synthesized as a single polypeptide chain that contains two distinct homologous lobes and a pro peptide, which is removed upon activation. In pepsinogen, the pro peptide precedes the N-terminal lobe (designated pep) and the C-terminal lobe (designated sin). Based on the three-dimensional structure of pepsinogen, we have designed a pepsinogen polypeptide with the internal rearrangement of domains from pro-pep-sin (native pepsinogen) to sin-pro-pep. The domain-rearranged zymogen also contains a 10-residue linker designed to connect sin and pro domains. Recombinant sin-pro-pep was synthesized in Escherichia coli, refolded from 8 M urea, and purified. Upon acidification, sin-pro-pep autoactivates to a two-chain enzyme. However, the emergence of activity is much slower than the conversion of the single-chain zymogen to a two-chain intermediate. In the activation of native pepsinogen and sin-pro-pep, the pro region is cleaved at two sites between residues 16P and 17P and 44P and 1 successively, and complete activation of sin-pro-pep requires an additional cleavage at a third site between residues 1P and 2P. In pepsinogen activation, the cleavage of the first site is rate limiting because the second site is cleaved more rapidly to generate activity. In the activation of sin-pro-pep, however, the second site is cleaved slower than the first, and cleavage of the third site is the rate limiting step.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Experiments were carried out on the effects of substrate or competitive inhibitor on the rate of appearance of N-terminal isoleucine residue of pepsin and peptides released from pepsinogen in its conversion to pepsin. Assumptions were made from these experiments, that an active site is initially formed in pepsinogen by acidification of its solution, and that peptide bond between 41-glutamyl and 42-isoleucyl residues locates in the juxtaposition to the active site forming an intramolecular enzyme-substrate complex. Thus, N-terminal tail of pepsinogen is released by a hydrolysis catalyzed by its own active site.

It was Indeed ascertained in this study that neither a small amount of pepsin which could be accompanied by pepsinogen preparation used contributes to the initial step of hydrolysis of pepsinogen nor pepsin formed accelerates the following activation process.

Therefore, it was concluded that the conversion of pepsinogen to pepsin is self-degrad-ation process.  相似文献   

4.
Early events of pepsinogen activation   总被引:1,自引:0,他引:1  
H E Auer  D M Glick 《Biochemistry》1984,23(12):2735-2739
Stopped-flow measurements both with native pig pepsinogen and with a fluorescent derivative, labeled near the carboxyl terminus with a toluidinylnaphthalenesulfonyl (TNS) group at Lys364, show rapid fluorescence changes following acidification. The rate constants observed by intrinsic fluorescence of the native zymogen are distinctly greater than those exhibited by the TNS derivative in the pH range examined. The rate constants for two early events in the activation of the derivative increase as the pH decreases from pH 3 to pH 2. The fluorescent intensities of these two processes also vary with pH. Because the ratios of these amplitudes fit the Henderson-Hasselbalch equation, it is concluded that the two processes represent concurrent events, rather than sequential ones. It is proposed that a protonation separates two forms of the zymogen. The conjugate acid undergoes the slower event, whereas the conjugate base, which predominates at pH 3, undergoes the faster event. It is proposed that both these pathways result in activation.  相似文献   

5.
Intramolecular pepsinogen activation is inhibited either by pepstatin, a potent pepsin inhibitor, or by purified globin from hemoglobin, a good pepsin substrate. Also, pepsinogen at pH 2 can be bound to a pepstatin-Sepharose column and recovered as native zymogen upon elution in pH 8 buffer. Kinetic studies of the globin inhibition of pepsinogen activation show that globin binds to a pepsinogen intermediate. This interaction gives rise to competitive inhibition of intramolecular pepsinogen activation. The evidence presented in this paper suggests that pepsinogen is converted rapidly upon acidification to the pepsinogen intermediate delta. In the absence of an inhibitor, the intermediate undergoes conformational change to bind the activation peptide portion of this same pepsinogen molecule in the active center to form an intramolecular enzyme-substrate complex (intermediate theta). This is followed by the intramolecular hydrolysis of the peptide bond between residues 44 and 45 of the pepsinogen molecule and the dissociation of the activation peptide from the pepsin. Intermediate delta apparently does not activate another pepsinogen molecule via an intermolecular process. Neither does intermediate delta hydrolyze globin substrate.  相似文献   

6.
Studies on the irreversible step of pepsinogen activation   总被引:1,自引:0,他引:1  
D M Glick  Y Shalitin  C R Hilt 《Biochemistry》1989,28(6):2626-2630
The bond cleavage step of pepsinogen activation has been investigated in a kinetic study in which the denatured products of short-term acidifications were separated on SDS-polyacrylamide gels and the peptide products were quantitated by densitometry. Although several peptide products were observed, under the conditions of the experiments (pH values between 2.0 and 2.8, 22 degrees C), the only one that was a product of an initial bond cleavage was the 44-residue peptide, which upon removal from pepsinogen yields pepsin. The rate constant for this bond cleavage is 0.015 s-1 at pH 2.4, which is the same as that at which the alkali-stable potential activity of pepsinogen had been found to convert to the alkali-labile activity of pepsin. When the conversion of zymogen to enzyme was followed by the change in fluorescence of adsorbed 6-(p-toluidinyl)naphthalene-2-sulfonate (TNS), the rate of change in TNS fluorescence was the same as the conversion to alkali lability. However, pepstatin blocked the bond cleavage of pepsinogen to pepsin, but it permitted the fluorescence change to proceed. In fact, it accelerated the apparent rate of change of TNS fluorescence by shifting the pKa of an essential conjugate acid from 1.7 to 2.6. The conversion to alkali lability, therefore, may be considered to be a composite of a relatively slow conformational change (at the measured rate), followed immediately by a relatively fast bond cleavage.  相似文献   

7.
The environment of the biotin binding site on avidin was investigated by determining the fluorescence enhancement of a series of fluorescent probes that are anilinonaphthalene sulfonic acid derivatives. Of the compounds tested, 2-anilinonaphthalene-6-sulfonic acid (2,6-ANS) exhibited the greatest enhancement under the conditions used (which would reflect both molar fluorescence enhancement and binding affinity) and exhibited more than 95% reversal upon addition of biotin. Thus, 2,6-ANS was chosen for more detailed characterization of the interaction with avidin. Only a single class of binding sites for 2,6-ANS was identified; the mean value for the Kd was 203 +/- 16 microM (X +/- 1 S.D.), and the molar ratio of 2,6-ANS binding sites to biotin binding sites was approx. 1. These results provide evidence that the biotin binding site and the 2,6-ANS binding site are at least partially overlapping, but the possibility that the probe binding site is altered by a conformational change induced by biotin binding cannot be excluded. At excitation = 328 nm and emission = 408 nm, the molar fluorescence of the bound probe was 6.8 +/- 1.0 microM-1 and that of the free probe was 0.061 +/- 0.008 microM-1 giving an enhancement ratio (molar fluorescence of bound probe/molar fluorescence of free probe) of 111 +/- 22. Upon binding, the wavelength of maximum fluorescence decreases. These findings also provide evidence that the fluorescence enhancement associated with the interaction of 2,6-ANS and avidin reflects the environment of the biotin binding site. The Kosower's Z factor, an empirical index of apolarity, was 82.1 for the 2,6-ANS binding site on avidin. This value reflects a degree of apolarity that is similar to apolar environments observed for substrate binding sites on several enzymes; although not the dominant factor, this environment may contribute to the strong binding of biotin.  相似文献   

8.
9.
Pig pepsinogen has been reacted with N-carboxymethylisatoic anhydride to form N-carboxymethyl-anthraniloyl-(CMA-) pepsinogen, derivatized at Lysp18, Lysp23, Lysp27, Lysp30, and Lys320. Conformational change associated with activation was detected by following energy transfer from tryptophan residues of the pepsin moiety, excited at 295 nm, to CMA groups, monitored by emission above 415 nm. Efficiency of this energy transfer is a measure of conformational change. For this zymogen derivative the change in efficiency occurs with a first order rate constant of 0.041 s-1 at pH 2.4, 22 degrees, which equals the rate at which, following acidification, alkali-stable potential activity becomes alkali-labile. For the native zymogen the rate of this conversion had been shown to be identical to the rate of cleavage of the scissile bond of pepsinogen. Therefore, the correspondence in this derivative of the rates of conversion to alkali lability and change in energy transfer demonstrates that a conformational change accompanies the peptide bond cleavage of activation.  相似文献   

10.
The complete amino acid sequence of monkey progastricsin   总被引:1,自引:0,他引:1  
The complete amino acid sequence of progastricsin from the Japanese monkey (Macaca fuscata) was determined. Progastricsin is composed of 374 residues, including the gastricsin moiety of 331 residues and the activation segment of 43 residues. Upon activation under acidic conditions, progastricsin was converted to gastricsin via the intermediate protein species. NH2-terminal sequence determination of these protein species enabled us to deduce the NH2-terminal 78-residue sequence of progastricsin, including the 43-residue activation segment. The complete sequence of the gastricsin moiety was determined using peptide fragments obtained by several chemical and enzymatic cleavages. The molecular weight of progastricsin was determined to be 40,785. As compared with pepsinogen A of the same monkey species, deletion of 4 residues and insertion of 5 residues were observed. Although monkey progastricsin and pepsinogen A have highly homologous sequences around the two active site aspartyl residues, the homology between these proteins is rather small (49% identity). This indicates that progastricsin diverged from pepsinogen A in the early phase of the evolution of gastric aspartyl proteinases.  相似文献   

11.
The activation process of pepsinogen was analyzed by a combination of computer simulation and experiment. In order to investigate in detail the behavior of the basic schemes proposed in the previous study, further computer simulations were conducted. Some experiments were performed based on the information obtained. The changes in the UV difference spectrum in the early stage was measured by the stopped-flow technique and the conversion of pepsinogen to pepsin [EC 3.4.23.1] was followed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Furthermore, on the basis of the experimental results, the most reasonable scheme was selected and modified. As a result, a scheme for the activation process of pepsinogen was obtained (Scheme 8). On the basis of the above analyses, it was assumed that the first step and the third step are pH-dependent based on the change in the UV spectrum, that the second step is a nonlinear reaction containing a looped reaction with a dimeric intermediate (in this step, peptide fragments are released and pepsinogen is converted to a pepsin-like molecule), and that the third step is an equilibrium reaction involving proton binding.  相似文献   

12.
Monkey pepsinogen A, monkey progastricsin, and porcine pepsinogen A were activated in the presence of two different protein substrates, namely, reduced and carboxymethylated lysozyme and hemoglobin. In each case, an extensive delay in activation was observed. The intermolecular activation reaction required for the generation of pepsin or gastricsin was strongly inhibited and this inhibition was essentially responsible for the delay. However, the intramolecular reaction required for the generation of the intermediate forms of the proenzymes was scarcely affected. The delay was longer at pH 3.0 than at pH 2.0. Irrespective of the delay in activation of pepsinogen, the digestion of substrates proceeded rapidly, evidence of the significant proteolytic activity of pepsinogen itself. Kinetic experiments demonstrated that pepsinogen changed from an enzymatically inactive species to an active species before the release of the activation segment. The proteolytic activity of the active pepsinogen was highest at pH 2.0, at 37 degrees C and the activity under these conditions was comparable to that of pepsin.  相似文献   

13.
Formation of reactive oxygen species in rat brain synaptosomes upon acidification of the incubation medium has been studied using a fluorescent probe DCFDA. It is shown that lowering the pH from 7.4 to 7.0 and 6.0 leads to a progressive increase in fluorescence that is indicative of oxidative stress. The effect is observed regardless of the presence of calcium ions in the incubation medium. In experiments without synaptosomes, acidification of the medium causes quenching of the fluorescence of DCF (pre-oxidized dye). This testifies that the fluorescence rise in synaptosomes is indeed associated with enhanced production of reactive oxygen species. Thus, even a small reduction of external pH to 7.0 is sufficient to cause oxidative stress in brain synaptosomes.  相似文献   

14.
Ca2+ binding to fast skeletal muscle troponin C reincorporated into troponin C-depleted (CDTA-treated) myofibrils has been measured directly by using 45Ca and indirectly by using a fluorescent probe. Direct Ca2(+)-binding measurements have shown that the Ca2+ affinity of the low-affinity sites is enhanced in the absence of ATP and conversely reduced when myosin is selectively extracted from myofibrils, compared to the Ca2+ affinity in the presence of ATP. Fluorescence intensity changes of a dansylaziridine label at the Met-25 residue of troponin C have shown the same Ca2(+)-sensitivity whether or not ATP is present, while much lower Ca2(+)-sensitivity is seen in the myosin-extracted myofibrils. Since the Met-25 residue is in the amino terminal side alpha-helix of Ca2(+)-binding site I and far from Ca2(+)-binding site II in the primary structure, Ca2+ binding to site II has been evaluated by assuming that the fluorescence change monitors Ca2+ binding to site I alone. Ca2+ binding to site II thus estimated has shown high positive cooperativity only in the presence of ATP and has been found to be nearly proportional to the activation of myofibrillar ATPase, suggesting that Ca2(+)-binding site II is directly involved in the activation of myofibrillar ATPase activity. On the other hand, Ca2(+)-binding site I has been suggested to regulate the interaction of weakly binding cross-bridges with the thin filament, since the fluorescence change in the presence of ATP is saturated at the free Ca2+ concentration required for the activation of myofibrillar ATPase.  相似文献   

15.
16.
Pepsinogen and Pepsin   总被引:2,自引:0,他引:2       下载免费PDF全文
Evidence relating to the structure and properties of swine pepsinogen and pepsin has been reviewed and used to suggest a tentative two dimensional picture of the skeleton of these two proteins. When pepsinogen, a folded single peptide chain, is converted to pepsin, there is a profound change in the physical and chemical properties of the protein. In an as yet unknown manner, except that it is initiated by a peptic cleavage of the protein chain, a single enzymic site is formed. This site is made up, quite probably, of the secondary carboxyl group of glutamic acid or of aspartic acid and a tyrosine phenol group in close proximity so that they can form hydrogen or hydrophobic bonds with the substrate in some unique manner that permits hydrolysis to occur at an accelerated rate.  相似文献   

17.
Derivatives of the fluorescent probes fluorescein and rhodamine specifically and covalently modify the highly reactive thiol (SH1) of myosin subfragment 1 (S1). Both probes develop circular dichroism (CD) upon modification of SH1 at the visible absorption band of the chromophore. A model system of chiral complexing agents (aromatic chiral amines) interacting with fluorescein in solvent develops a CD signal that mimics that produced by S1. The model system suggests that a specific interaction of the probe with an aromatic chiral residue in the SH1 binding pocket induces the CD signal. Several other spectroscopic signals, including absorption and fluorescence intensity and anisotropy, characterize the fluorescein or rhodamine binding to SH1. A coupled dipole method is adapted to interpret these spectroscopic signals in terms of the probe-S1 complex conformation. The computation of the orientation of the principal hydrodynamic frame (PHF) of S1 from its crystallographic -carbon backbone structure permits the known orientation of the probe in the PHF of S1 to further constrain the conformation of the probe-S1 complex. The coupled dipole interpretation of spectroscopic data combined with constraints relating the probe dipole orientation to the PHF of S1 determines the conformation of the probe-S1 complex. The methods developed here are applied to the spectroscopic signals from fluorescein or rhodamine in the SH1 binding site of S1 to obtain an atomic resolution model of the probe-S1 conformation [Ajtai and Burghardt, Biochemistry, 34 (1995) 15943–15952].  相似文献   

18.
The entire pepsinogen C (PGC) coding sequence was determined by analysis of a series of five overlapping cDNA clones identified in a library constructed from human gastric mucosa poly(A+) RNA. A partial cDNA clone was initially identified using a 256-fold degenerate oligonucleotide probe for amino acid residues 4-12 of pepsin C, and subsequently 4 additional clones were identified upon rescreening with a probe complementary to the 5' region of the original cDNA clone. Northern analysis of gastric mucosa poly(A+) RNA with a PGC cDNA probe revealed an mRNA 1.5-kilobase species that was indistinguishable from that detected with a human pepsinogen A (PGA) cDNA probe. In contrast, the PGC and PGA cDNA probes detected distinct genomic restriction fragments indicating there was no detectable cross-hybridization under high stringency conditions. The PGC gene was localized to human chromosome 6 by analysis of a panel of human x mouse somatic cell hybrids. The regions containing the active site aspartyl groups of PGC are conserved in relationship to several other aspartic proteinases. We propose that the absence of detectable immunologic cross-reactivity between the two groups of human pepsinogens, A and C, results from divergent evolution of sequences located on the surface of the zymogens in contrast to the strongly conserved active site regions located within the binding cleft of the enzymes that are inaccessible for antigenic recognition.  相似文献   

19.
Identification problems restrict quantitative ecological research on specific nanoflagellates. Identification by specific oligonucleotide probes permits use of flow cytometry for enumeration and measurement of size of nanoflagellates in statistically meaningful samples. Flow cytometry also permits measurement of intensity of probe binding by cells. Five fluorescent probes targeted to different regions of the small subunit rRNA of the common marine flagellate Paraphysomonas vestita all hybridized with cells of this flagellate. Cells fixed with trichloroacetic acid gave detectable signals at a probe concentration of 15 aM and specific fluorescence increased almost linearly to 1.5 fM, but at higher concentrations nonspecific binding increased sharply. Three flagellates, P. vestita, Paraphysomonas imperforata, and Pteridomonas danica, all bound a general eukaryotic probe approximately in proportion to their cell size, but the specific P. vestita probe gave 14 times more fluorescence with P. vestita than with either of the other flagellates. Cell fluorescence increased during the early growth of a batch culture and decreased toward the stationary phase; cell size changed in a comparable manner. Cell fluorescence intensity may allow inferences about growth rate, but whether fluorescence (assumed to reflect ribosome number) merely correlates with cell biomass or changes in a more complex manner remains unresolved.  相似文献   

20.
Conformational changes in the beta-subunit of the bovine brain Ca2+-binding protein S100b (S100-beta) accompanying Ca2+ binding were investigated by analysis of the spectroscopic properties of the single tyrosine residue (Tyr17 beta) and flow-dialysis binding experiments. S100-beta binds Ca2+ sequentially at two sites to change the conformation of the protein. The first Ca2+ ion binds to site II beta, a typical Ca2+-binding site in the C-terminal region, and it does not significantly perturb the proximal environment of Tyr17 beta. After the first site is occupied, another Ca2+ ion binds to the N-terminal Ca2+-binding site, I beta, and strengthens a hydrogen bond between Tyr17 beta and a neighbouring carboxylate acceptor group, which results in a large increase in the Tyr17 beta fluorescence spectrum half-width and a positive absorption and c.d. signal between 290 and 275 nm. Ca2+ binding to the S100b.Zn2+6 complex, studied by flow-dialysis and fluorescence measurements showed that, although Zn2+ ions increase the affinity of S100b protein for Ca2+, the Ca2+-binding sequence was not changed. Tb3+ (terbium ion) binding studies on the S100b.Zn2+6 complex proved that Tb3+ antagonizes only Ca2+ binding site II beta and confirmed the sequential occupation of Ca2+-binding sites on the S100b.Zn2+6 complex.  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