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 0 毫秒
1.
    
Endothiapepsin is a typical member of the aspartic proteinase family. The catalytic mechanism of this family is attributed to two conserved catalytic aspartate residues, which coordinate the hydrolysis of a peptide bond. An oligopeptide inhibitor (IC50 = 0.62 µM) based on a reduced‐bond transition‐state inhibitor of mucorpepsin was co‐crystallized with endothiapepsin and the crystal structure of the enzyme–inhibitor complex was determined at 1.35 Å resolution. A total of 12 hydrogen bonds between the inhibitor and the active‐site residues were identified. The resulting structure demonstrates a number of novel subsite interactions in the active‐site cleft.  相似文献   

2.
A Gustchina  I T Weber 《Proteins》1991,10(4):325-339
The different isolates available for HIV-1 and HIV-2 were compared for the region of the protease (PR) sequence, and the variations in amino acids were analyzed with respect to the crystal structure of HIV-1 PR with inhibitor. Based on the extensive homology (39 identical out of 99 residues), models were built of the HIV-2 PR complexed with two different aspartic protease inhibitors, acetylpepstatin and a renin inhibitor, H-261. Comparison of the HIV-1 PR crystal structure and the HIV-2 PR model structure and the analysis of the changes found in different isolates showed that correlated substitutions occur in the hydrophobic interior of the molecule and at surface residues involved in ionic or hydrogen bond interactions. The substrate binding residues of HIV-1 and HIV-2 PRs show conservative substitutions of four residues. The difference in affinity of HIV-1 and HIV-2 PRs for the two inhibitors appears to be due in part to the change of Val 32 in HIV-1 PR to Ile in HIV-2 PR.  相似文献   

3.
    
Plasmepsin (Plm) has been identified as an important target for the development of new antimalarial drugs, since its inhibition leads to the starvation of Plasmodium falciparum. A series of substrate-based dipeptide-type Plm II inhibitors containing the hydroxymethylcarbonyl isostere as a transition-state mimic were synthesized. The general design principle was provision of a conformationally restrained hydroxyl group (corresponding to the set residue at the P2' position in native substrates) and a bulky unit to fit the S2' pocket.  相似文献   

4.
The crystal structure of Bla g 2 was solved in order to investigate the structural basis for the allergenic properties of this unusual protein. This is the first structure of an aspartic protease in which conserved glycine residues, in two canonical DTG triads, are substituted by different amino acid residues. Another unprecedented feature revealed by the structure is the single phenylalanine residue insertion on the tip of the flap, with the side-chain occupying the S1 binding pocket. This and other important amino acid substitutions in the active site region of Bla g 2 modify the interactions in the vicinity of the catalytic aspartate residues, increasing the distance between them to approximately 4A and establishing unique direct contacts between the flap and the catalytic residues. We attribute the absence of substantial catalytic activity in Bla g 2 to these unusual features of the active site. Five disulfide bridges and a Zn-binding site confer stability to the protein, which may contribute to sensitization at lower levels of exposure than other allergens.  相似文献   

5.
Lysosomal proteases are actively involved into pathogenesis of malignant tumors. Impairments in the interaction between proteases and their inhibitors are implicated in the processes of tumor invasion and metastasis. Among proteases associated with malignant growth, cysteine cathepsins B and L and aspartic cathepsin D are considered to play the major role in the tumor development. The present study was designed to investigate the activity of cathepsins B, L, and D during the development and treatment of murine experimental leukemias and to determine correlation between these proteases and course of pathological process as well as efficiency of the chemotherapeutic treatment. P-388 leukemia was characterized by a more aggressive development and unfavorable prognosis than L1210/1 leukemia. In mice with P-388 leukemia the activity of lysosomal cathepsins B, D, and L in the tumor tissue, liver and spleen, as well as the activity of cathepsins B and L in serum were lower than activities of these enzymes in mice with L1210/1 leukemia. Changes in the activity of cathepsins in liver and spleen of leukemic mice reflected a level of aggressiveness of the tumor development and invasion of these organs with tumor cells. Treatment of these experimental leukemias resulted in the increase of cathepsin B, L and D activity in the tumor tissue, liver, spleen and the increase in cathepsin B and L activity in serum. The highest protease activity was detected in the groups of mice characterized by the highest inhibition of the tumor growth. These data demonstrate that lysosomal proteases are involved in the progression of murine experimental leukemias and elimination of tumor cells in the result of treatment. Thus, determination of the activity of cysteine and aspartic proteases can be used for evaluation of cancer malignancy, tumor sensitivity for chemotherapy and efficiency of treatment.  相似文献   

6.
The crystal structures, dimer stabilities, and kinetics have been analyzed for wild-type human immunodeficiency virus type 1 (HIV-1) protease (PR) and resistant mutants PR(L24I), PR(I50V), and PR(G73S) to gain insight into the molecular basis of drug resistance. The mutations lie in different structural regions. Mutation I50V alters a residue in the flexible flap that interacts with the inhibitor, L24I alters a residue adjacent to the catalytic Asp25, and G73S lies at the protein surface far from the inhibitor-binding site. PR(L24I) and PR(I50V), showed a 4% and 18% lower k(cat)/K(m), respectively, relative to PR. The relative k(cat)/K(m) of PR(G73S) varied from 14% to 400% when assayed using different substrates. Inhibition constants (K(i)) of the antiviral drug indinavir for the reaction catalyzed by the mutant enzymes were about threefold and 50-fold higher for PR(L24I) and PR(I50V), respectively, relative to PR and PR(G73S). The dimer dissociation constant (K(d)) was estimated to be approximately 20 nM for both PR(L24I) and PR(I50V), and below 5 nM for PR(G73S) and PR. Crystal structures of the mutants PR(L24I), PR(I50V) and PR(G73S) were determined in complexes with indinavir, or the p2/NC substrate analog at resolutions of 1.10-1.50 Angstrom. Each mutant revealed distinct structural changes relative to PR. The mutated residues in PR(L24I) and PR(I50V) had reduced intersubunit contacts, consistent with the increased K(d) for dimer dissociation. Relative to PR, PR(I50V) had fewer interactions of Val50 with inhibitors, in agreement with the dramatically increased K(i). The distal mutation G73S introduced new hydrogen bond interactions that can transmit changes to the substrate-binding site and alter catalytic activity. Therefore, the structural alterations observed for drug-resistant mutations were in agreement with kinetic and stability changes.  相似文献   

7.
    
The substrate specificity of porcine pepsin has been altered by site-directed mutagenesis in an attempt to selectively cleave bovine hide collagen at only a few sites, similar to cathepsin D, for the production of high quality gelatin. Kinetic parameters were determined using chromogenic peptide substrates based on the sequence Lys-Pro-Xaa-Yaa-Phe*Nph-Arg-Leu (where Xaa is Ile or Pro, Yaa is Glu. Leu, Gln or Lys, Nph is p-nitrophenylalanine, and * is the site of cleavage). Substitution of Thr222 and Glu287 within the S2 subsite of pepsin by Val and Met, respectively, produced a double mutant with a two- to fourfold higher kcat/Km, compared with wild-type pepsin, for the chromogenic peptides with residues Leu, Gln, and Glu at position P2 (Yaa). The results suggest that the functional group of the P2 side chain may be exposed to solvent, while the aliphatic portion interacts with hydrophobic residues comprising S2. Wild-type pepsin cleaved a peptide corresponding to the carboxy-terminal telopeptide region of bovine type I collagen alpha1 chain, SGGYDLSFLPQPPQE, predominantly at three sites (Asp-Leu, Leu-Ser, and Phe-Leu) and at a significantly lower rate at Ser-Phe. However, Thr222Val/Glu287Met cleaved site Ser-Phe at a rate 20-fold higher than the wild-type. Significantly, enzymes containing the double substitution Phe111Thr/Leu112Phe cleaved this peptide predominantly at one site Leu-Ser (similar to cathepsin D) and at a rate 23-fold higher than the wild-type. These mutants can potentially enhance the rate of solubilization of bovine hide collagen under conditions mild enough to maintain the triple helix structure and hence minimize the rate of subsequent denaturation and proteolytic cleavage.  相似文献   

8.
Evolution in the structure and function of aspartic proteases   总被引:22,自引:0,他引:22  
Aspartic proteases (EC3.4.23) are a group of proteolytic enzymes of the pepsin family that share the same catalytic apparatus and usually function in acid solutions. This latter aspect limits the function of aspartic proteases to some specific locations in different organisms; thus the occurrence of aspartic proteases is less abundant than other groups of proteases, such as serine proteases. The best known sources of aspartic proteases are stomach (for pepsin, gastricsin, and chymosin), lysosomes (for cathepsins D and E), kidney (for renin), yeast granules, and fungi (for secreted proteases such as rhizopuspepsin, penicillopepsin, and endothiapepsin). These aspartic proteases have been extensively studied for their structure and function relationships and have been the topics of several reviews or monographs (Tang: Acid Proteases, Structure, Function and Biology. New York: Plenum Press, 1977; Tang: J Mol Cell Biochem 26:93-109, 1979; Kostka: Aspartic Proteinases and Their Inhibitors. Berlin: Walter de Gruyter, 1985). All mammalian aspartic proteases are synthesized as zymogens and are subsequently activated to active proteases. Although a zymogen for a fungal aspartic protease has not been found, the cDNA structure of rhizopuspepsin suggests the presence of a "pro" enzyme (Wong et al: Fed Proc 44:2725, 1985). It is probable that other fungal aspartic proteases are also synthesized as zymogens. It is the aim of this article to summarize the major models of structure-function relationships of aspartic proteases and their zymogens with emphasis on more recent findings. Attempts will also be made to relate these models to other aspartic proteases.  相似文献   

9.
    
The crystal structures of two constructs of RC1339/APRc from Rickettsia conorii, consisting of either residues 105–231 or 110–231 followed by a His tag, have been determined in three different crystal forms. As predicted, the fold of a monomer of APRc resembles one‐half of the mandatory homodimer of retroviral pepsin‐like aspartic proteases (retropepsins), but the quaternary structure of the dimer of APRc differs from that of the canonical retropepsins. The observed dimer is most likely an artifact of the expression and/or crystallization conditions since it cannot support the previously reported enzymatic activity of this bacterial aspartic protease. However, the fold of the core of each monomer is very closely related to the fold of retropepsins from a variety of retroviruses and to a single domain of pepsin‐like eukaryotic enzymes, and may represent a putative common ancestor of monomeric and dimeric aspartic proteases.  相似文献   

10.
Andreeva  N. S.  Popov  M. E. 《Molecular Biology》2002,36(5):760-765
In addition to previous studies, 30 crystal structures of retroviral proteases corresponding to the highest resolution were inspected to analyze the interactions of the active carboxyls with surroundings groups. The outer oxygens of the active carboxyls in retroviral enzymes form contacts only with the water molecule participating in catalysis. This is an important difference between retroviral proteases and pepsin-like enzymes, which form a net of hydrogen bonds of these outer oxygens with residues neighboring the catalytic site in 3D structures. At the same time, it was found that in all aspartic proteases the inner oxygens of the active carboxyls are also involved in the chain of interactions through peptide groups Thr–Gly adjacent to the active residues. Polarization of these peptide groups influences the donor–acceptor properties of the active carboxyls. The found chain of interactions is more extensive in retroviral than in pepsin-like proteases; however, its main part is conserved for the whole class of these enzymes. Some implications of the role of these interactions are discussed.  相似文献   

11.
Summary In facilitated transport systems the carrier reorientation step is shown to be largely independent of the forces of interaction between the substrate and the carrier site, whereas in coupled systems (obligatory exchange or cotransport) reorientation proceeds at the expense of the binding force developed in the transition state. In consequence, the expression of substrate specificity is expected to differ in the two systems. In the facilitated transport of analogs no larger than the normal substrate, the affinity but not the maximum rate of transport can vary widely; with larger analogs, both the affinity and rale can vary if steric constraints are more severe in the translocation step than in binding. In coupled transport, by contrast, the translocation step can be highly sensitive to the structure of the substrate, and binding much less sensitive. The theory agrees with published observations on facilitated systems for choline and glucose in erythrocytes, as well as on Na+-coupled systems for the same substrates in other cells. The following mechanism, which could account for the behavior, is proposed. In facilitated systems, the transport site fits the substrate closely and retains its shape as the carrier undergoes reorientation. In coupled systems, the site is initially looser, but during carrier reorientation it contracts around the substrate. In both systems, the carrier encloses the substrate during the translocation step, though for a different reason: in coupled but not in facilitated systems the binding force enormously increases in the enclosed state, through a chelation effect. In both systems, steric interference with enclosure retards the translocation of bulky substrate analogs.  相似文献   

12.
Summary An analysis of transition-state models for exchange-only transport shows that substrate binding forces, carrier conformational changes, and coupled substrate flow are interrelated. For a system to catalyze exchange but not net transport, addition of the substrate must convert the carrier from an immobile to a mobile form. The reduction in the energy barrier to movement is necessarily paid for out of the intrinsic binding energy between the substrate and the transport site, and is dependent on the formation of two different types of complex: a loose complex initially and a tight complex in the transition state in carrier movement. Hence the site should at first be incompletely organized for optimal binding but, following a conformational change, complementary to the substrate structure in the transition state. The conformational change, which may involve the whole protein, would be induced by cooperative interactions between the substrate and several groups within the site, involving a chelate effect. The tightness of coupling, i.e., the ratio of exchange to net transport, is directly proportional to the increased binding energy in the transition state, a relationship which allows the virtual substrate dissociation constant in the transition state to be calculated from experimental rate and half-saturation constants. Because the transition state is present in minute amount, strong bonding here does not enhance the substrate's affinity, and specificity may, therefore, be expressed in maximum exchange rates alone. However, where substrates largely convert the carrier to a transport intermediate whose mobility is the same with all substrates, specificity is also expressed in affinity. Hence the expression of substrate specificity provides evidence on the translocation mechanism.  相似文献   

13.
The characteristic mode for the herbicidal action of 2-amino-3-chloro-1,4-naphthoquinone (ACN) was investigated by using autotrophic green microalgae (Scenedesmus acutus). The effects of ACN on the growth, chlorophyll content, protoporphyrin-IX accumulation, and ethane production in Scenedesmus cells were measured in comparison with three reference herbicides, i.e., ioxynil, dinoseb, and chlorophthalim. ACN appeared to have a different mechanism for action from these three herbicides. S. acutus cells grown with ACN produced ethane, but the mechanism for its production is considered to have been different from that of chlorophthalim. ACN may inhibit porphyrin biosynthesis at a different stage from that by chlorophthalim.  相似文献   

14.
A brief account is given of the specific interactions of some amino acid residues in aspartic proteases of both higher organisms and retroviruses that determine important protease properties: the anomalously low isoelectric point of pepsin and its stability at pH close to 1; the ability of one of the carboxyl groups in the active site of proteases of higher organisms to retain a charged state at any pH value; and the protonated state of another carboxyl, which is necessary for enzymatic activity. It is also explained how such states can be induced in retroviral proteases.  相似文献   

15.
Allan Beveridge 《Proteins》1996,24(3):322-334
We have performed ab initio Hartree-Fock self-consistent field calculations on the active site of endothiapepsin. The active site was modeled as a formic acid/formate anion moiety (representing the catalytic aspartates, Asp-32 and -215) and a bound water molecule. Residues Gly-34, Ser-35, Gly-217, and Thr-218, which all form hydrogen bonds to the active site, were modeled using formamide and methanol molecules. The water molecule, which is generally believed to function as the attacking nucleophile in catalysis, was allowed to bind to the active site in four distinct configurations. The geometry of each configuration was optimized using two basis sets (4-31G and 4-31G*). The results indicate that in the native enzyme the nucleophilic water is bound in a catalytically inert configuration. However, by rotating the carboxyl group of Asp-32 by about 90° the water molecule can be reorientated to attack the scissile bond of the substrate. A model of the bound enzyme-substrate complex was constructed from the crystal structure of a difluorostatone inhibitor complexed with endothiapepsin. This model suggests that the substrate itself initiates the reorientation of the nucleophilic water immediately prior to catalysis by forcing the carboxyl group of Asp-32 to rotate. The theoretical results predict that the active site of endothiapepsin undergoes a large distortion during substrate binding and this observation has been used to explain some of the kinetics results which have been reported for mutant aspartic proteinases.  相似文献   

16.
    
Plasmepsin 4 from the malarial parasite Plasmodium malariae (PmPM4) is a member of the plasmepsins (Plasmodium pepsins), a subfamily of the pepsin‐like aspartic proteases whose ortholog in the malarial parasite P. falciparum is involved in hemoglobin digestion in its digestive vacuole. Crystals of PmPM4 in complex with the small‐molecule inhibitor AG1776 have been grown from a precipitant of 15% PEG 4000 and 200 mM ammonium sulfate in 100 mM sodium acetate pH 4.5. X‐ray diffraction data were collected on a Rigaku rotating‐anode generator from a single crystal under cryoconditions, with a maximal useful diffraction pattern to 3.3 Å resolution. The crystals are shown to be orthorhombic and have been assigned to space group P21212, with unit‐cell parameters a = 95.88, b = 112.58, c = 90.40 Å and a scaling Rsym of 0.104 for 14 334 unique reflections. Packing consideration and self‐rotation function results indicate that there are two molecules per asymmetric unit. It is expected that in the near future the structure of PmPM4 will be obtained using molecular‐replacement methods, obtaining phases from previously determined plasmepsin structures. Elucidation of the structure of PmPM4 in complex with inhibitors may be paramount to producing new antimalarial therapeutic agents.  相似文献   

17.
    
The sequence–structure–function paradigm of proteins has been revolutionized by the discovery of intrinsically disordered proteins (IDPs) or intrinsically disordered regions (IDRs). In contrast to traditional ordered proteins, IDPs/IDRs are unstructured under physiological conditions. The absence of well‐defined three‐dimensional structures in the free state of IDPs/IDRs is fundamental to their function. Folding upon binding is an important mode of molecular recognition for IDPs/IDRs. While great efforts have been devoted to investigating the complex structures and binding kinetics and affinities, our knowledge on the binding mechanisms of IDPs/IDRs remains very limited. Here, we review recent advances on the binding mechanisms of IDPs/IDRs. The structures and kinetic parameters of IDPs/IDRs can vary greatly, and the binding mechanisms can be highly dependent on the structural properties of IDPs/IDRs. IDPs/IDRs can employ various combinations of conformational selection and induced fit in a binding process, which can be templated by the target and/or encoded by the IDP/IDR. Further studies should provide deeper insights into the molecular recognition of IDPs/IDRs and enable the rational design of IDP/IDR binding mechanisms in the future.  相似文献   

18.
19.
    
The human immunodeficiency virus (HIV) codes for an aspartic protease known to be essential for retroviral maturation and replication. HIV protease is formed from two identical 99 amino acid peptides. We synthesized [(NHCH2CH2—S—CH2CO)51–52, Ala67,95]HIV‐1 protease using the thioether chemical ligation method, and then prepared the [(NHCH2CH2—S—CH2CO)51–52, Ala67,95, Cys98]HIV‐1 protease dimer analogue covalently linked by a disulfide bridge. These HIV‐1 protease analogues effectively cleaved the Tyr–Phe‐type substrate, but had weak affinity to the Tyr–Pro‐type substrate. Consequently, the molecular recognition of the protease analogues differs from that of the wild‐type enzyme. Based on the substrate transition state, we designed and synthesized a novel class of HIV protease inhibitors containing an unnatural amino acid, (2S, 3S)‐3‐amino‐2‐hydroxy‐4‐phenylbutyric acid, named allophenylnorstatine, with a hydroxymethylcarbonyl (HMC) isostere. The stereochemistry of the hydroxyl group was significant for the enzyme inhibition and the HMC group interacted excellently with the aspartic acid carboxyl groups of HIV protease active site in the essentially same hydrogen‐bonding mode as the transition state. Small dipeptide‐based HIV protease inhibitors containing the HMC isostere were studied as advantageous compounds. Among them, a dipeptide‐based HIV protease inhibitor, KNI‐577, exhibited potent antiviral activities, low cytotoxicity, and good pharmacokinetic properties. © 1999 John Wiley & Sons, Inc. Biopoly 51: 59–68, 1999  相似文献   

20.
yapsin蛋白酶家族研究进展   总被引:1,自引:0,他引:1  
yapsin蛋白酶是一类糖基磷脂酰肌醇(GPI)锚定的天冬氨酸蛋白酶,能特异性地切割底物中的单或双碱性氨基酸残基位点,起着加工前肽的作用。近年来研究发现,工程菌表达某些外源蛋白时发生的降解现象与yapsin蛋白酶有紧密的关系。概述了yapsin蛋白酶家族的命名、结构、定位、酶原激活、基因表达及底物特异性和功能。  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