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 234 毫秒
1.
HTLV-I is an oncogenic retrovirus that is associated with adult T-cell leukemia. HTLV-I protease and HTLV-I protease fused to a deca-histidine containing leader peptide (His-protease) have been cloned, expressed, and purified. The refolded proteases were active and exhibited nearly identical enzymatic activities. To begin to characterize the specificity of HTLV-I, we measured protease cleavage of peptide substrates and inhibition by protease inhibitors. HTLV-I protease cleavage of a peptide representing the HTLV-I retroviral processing site P19/24 (APQVLPVMHPHG) yielded Km and kcat values of 470 microM and 0.184 s-1 while cleavage of a peptide representing the processing site P24/15 (KTKVLVVQPK) yielded Km and kcat values of 310 microM and 0.0060 s-1. When the P1' proline of P19/24 was replaced with p-nitro-phenylalanine (Nph), the ability of HTLV-I protease to cleave the substrate (APQVLNphVMHPL) was improved. Inhibition of HTLV-I protease and His-protease by a series of protease inhibitors was also tested. It was found that the Ki values for inhibition of HTLV-I protease and His-protease by a series of pepsin inhibitors ranged from 7 nM to 10 microM, while the Ki values of a series of HIV-1 protease inhibitors ranged from 6 nM to 127 microM. In comparison, the Ki values for inhibition of pepsin by the pepsin inhibitors ranged from 0.72 to 19.2 nM, and the Ki values for inhibition of HIV-1 protease by the HIV protease inhibitors ranged from 0.24 nM to 1.0 microM. The data suggested that the substrate binding site of HTLV-I protease is different from the substrate binding sites of pepsin and HIV-1 protease, and that currently employed HIV-1 protease inhibitors would not be effective for the treatment of HTLV-I infections.  相似文献   

2.
The design, synthesis, and molecular modeling studies of a novel series of azacyclic ureas, which are inhibitors of human immunodeficiency virus type 1 (HIV-1) protease that incorporate different ligands for the S1', S2, and S2' substrate-binding sites of HIV-1 protease are described. The synthesis of this series is highly flexible in the sense that the P1', P2, and P2' residues of the inhibitors can be changed independently. Molecular modeling studies on the phenyl ring of the P2 and P2' ligand suggested incorporation of hydrogen-bonding donor/acceptor groups at the 3' and 4-positions of the phenyl ring should increase binding potency. This led to the discovery of compound 7f (A-98881), which possesses high potency in the HIV-1 protease inhibition assay and the in vitro MT-4 cell culture assay (Ki = approximately 5 pM and EC50 = 0.002 microM). This compares well with the symmetrical cyclic urea 1 pioneered at DuPont Merck.  相似文献   

3.
A study on the use of derivatized carbohydrates as C2-symmetric HIV-1 protease inhibitors has been undertaken. L-Mannaric acid (6) was bis-O-benzylated at C-2 and C-5 and subsequently coupled with amino acids and amines to give C2-symmetric products based on C-terminal duplication. Potent HIV protease inhibitors, 28 Ki = 0.4 nM and 43 Ki = 0.2 nM, have been discovered, and two synthetic methodologies have been developed, one whereby these inhibitors can be prepared in just three chemical steps from commercially available materials. A remarkable increase in potency going from IC50 = 5000 nM (23) to IC50 = 15 nM (28) was observed upon exchanging -COOMe for -CONHMe in the inhibitor, resulting in the net addition of one hydrogen bond interaction between each of the two -NH- groups and the HIV protease backbone (Gly 48/148). The X-ray crystal structures of 43 and of 48 have been determined (Figures 5 and 6), revealing the binding mode of these inhibitors which will aid further design.  相似文献   

4.
As long as the threat of human immunodeficiency virus (HIV) protease drug resistance still exists, there will be a need for more potent antiretroviral agents. We have therefore determined the crystal structures of HIV-1 protease in complex with six cyclic urea inhibitors: XK216, XK263, DMP323, DMP450, XV638, and SD146, in an attempt to identify 1) the key interactions responsible for their high potency and 2) new interactions that might improve their therapeutic benefit. The structures reveal that the preorganized, C2 symmetric scaffolds of the inhibitors are anchored in the active site of the protease by six hydrogen bonds and that their P1 and P2 substituents participate in extensive van der Waals interactions and hydrogen bonds. Because all of our inhibitors possess benzyl groups at P1 and P1', their relative binding affinities are modulated by the extent of their P2 interactions, e.g. XK216, the least potent inhibitor (Ki (inhibition constant) = 4.70 nM), possesses the smallest P2 and the lowest number of P2-S2 interactions; whereas SD146, the most potent inhibitor (Ki = 0.02 nM), contains a benzimidazolylbenzamide at P2 and participates in fourteen hydrogen bonds and approximately 200 van der Waals interactions. This analysis identifies the strongest interactions between the protease and the inhibitors, suggests ways to improve potency by building into the S2 subsite, and reveals how conformational changes and unique features of the viral protease increase the binding affinity of HIV protease inhibitors.  相似文献   

5.
Integration of a cDNA copy of the human immunodeficiency virus (HIV) genome is mediated by an HIV-1-encoded enzyme, integrase (IN), and is required for productive infection of CD4+ lymphocytes. It had been shown that 3,5-dicaffeoylquinic acid and two analogues were potent and selective inhibitors of HIV-1 IN in vitro. To determine whether the inhibition of IN by dicaffeoylquinic acids was limited to the 3,5 substitution, 3,4-, 4,5-, and 1,5-dicaffeoylquinic acids were tested for inhibition of HIV-1 replication in tissue culture and inhibition of HIV-1 IN in vitro. All of the dicaffeoylquinic acids were found to inhibit HIV-1 replication at concentrations ranging from 1 to 6 microM in T cell lines, whereas their toxic concentrations in the same cell lines were > 120 microM. In addition, the compounds inhibited HIV-1 IN in vitro at submicromolar concentrations. Molecular modeling of these ligands with the core catalytic domain of IN indicated an energetically favorable reaction, with the most potent inhibitors filling a groove within the predicted catalytic site of IN. The calculated change in internal free energy of the ligand/IN complex correlated with the ability of the compounds to inhibit HIV-1 IN in vitro. These results indicate that the dicaffeoylquinic acids as a class are potent and selective inhibitors of HIV-1 IN and form important lead compounds for HIV drug discovery.  相似文献   

6.
To gain greater understanding of the structural basis of human immunodeficiency virus (HIV) protease ligand specificity, we have crystallized and determined the structures of the HIV-1 protease (Val32Ile, Ile47Val, Val82Ile) triple mutant and simian immunodeficiency virus (SIV) protease in complex with SB203386, a tripeptide analogue inhibitor containing a C-terminal imidazole substituent as an amide bond isostere. SB203386 is a potent inhibitor of HIV-1 protease (Ki = 18 nM) but shows decreased inhibition of the HIV-1 protease (Val32Ile, Ile47Val, Val82Ile) triple mutant (Ki = 112 nM) and SIV protease (Ki = 960 nM). Although SB203386 binds in the active site cavity of the triple mutant in a similar fashion to its binding to the wild-type HIV-1 protease [Abdel-Meguid et al. (1994) Biochemistry 33, 11671], it binds to SIV protease in an unexpected mode showing two inhibitor molecules each binding to half of the active site. Comparison of these two structures and that of the wild-type HIV-1 protease bound to SB203386 reveals that HIV protease ligand specificity is imparted by residues outside of the catalytic pocket, which causes subtle changes in its shape. Furthermore, this work illustrates the importance of structural studies in order to understand the structure-activity relationship (SAR) between related enzymes.  相似文献   

7.
Resistance of HIV-1 to protease inhibitors has been associated with changes at residues Val82 and Ile84 of HIV-1 protease (HIV PR). Using both an enzyme assay with a peptide substrate and a cell-based infectivity assay, we examined the correlation between the inhibition constants for enzyme activity (Ki values) and viral replication (IC90 values) for 5 active site mutants and 19 protease inhibitors. Four of the five mutations studied (V82F, V82A, I84V, and V82F/I84V) had been identified as conferring resistance during in vitro selection using a protease inhibitor. The mutant protease genes were expressed in Escherichia coli for preparation of enzyme, and inserted into the HXB2 strain of HIV for test of antiviral activity. The inhibitors included saquinavir, indinavir, nelfinavir, 141W94, ritonavir (all in clinical use), and 14 cyclic ureas with a constant core structure and varying P2, P2' and P3, P3' groups. The single mutations V82F and I84V caused changes with various inhibitors ranging from 0.3- to 86-fold in Ki and from 0.1- to 11-fold in IC90. Much larger changes compared to wild type were observed for the double mutation V82F/I84V both for Ki (10-2000-fold) and for IC90 (0.7-377-fold). However, there were low correlations (r2 = 0.017-0.53) between the mutant/wild-type ratio of Ki values (enzyme resistance) and the mutant/wild-type ratio of viral IC90 values (antiviral resistance) for each of the HIV proteases and the viruses containing the identical enzyme. Assessing enzyme resistance by "vitality values", which adjust the Ki values with the catalytic efficiencies (kcat/Km), caused no significant improvement in the correlation with antiviral resistance. Therefore, our data suggest that measurements of enzyme inhibition with mutant proteases may be poorly predictive of the antiviral effect in resistant viruses even when mutations are restricted to the protease gene.  相似文献   

8.
The activity of three human immunodeficiency virus (HIV) protease inhibitors was investigated in human primary monocytes/macrophages (M/M) chronically infected by HIV-1. Saquinavir, KNI-272, and ritonavir inhibited the replication of HIV-1 in vitro, with EC50s of approximately 0.5-3.3 microM. However, only partial inhibition was achievable, even at the highest concentrations tested. Also, the activity of these drugs in chronically infected M/M was approximately 7- to 26-fold lower than in acutely infected M/M and approximately 2- to 10-fold lower than in chronically infected H9 lymphocytes. When protease inhibitors were removed from cultures of chronically infected M/M, production of virus rapidly returned to the levels found in untreated M/M. Therefore, relatively high concentrations of protease inhibitors are required to suppress HIV-1 production in chronically infected macrophages, and such cells may be a vulnerable point for the escape of virus in patients taking these drugs.  相似文献   

9.
10.
Comparison of the high-resolution X-ray structures of the native HIV-1 protease and its complexes with the inhibitors suggested that the enzyme flaps are flexible. The movement at the tip of the flaps could be as large as 7 A. On the basis of this observation, cyclic cyanoguanidines have been designed, synthesized, and evaluated as HIV-1 protease (PR) inhibitors. Cyclic cyanoguanidines were found to be very potent inhibitors of HIV-1 protease. The choice of cyclic cyanoguanidines over cyclic guanidines was based on the reduced basicity of the former. X-ray structure studies of the HIV PR complex with cyclic cyanoguanidine demonstrated that in analogy to cyclic urea, cyclic cyanoguanidines also displace the unique structural water molecule. The structure-activity relationship of the cyclic cyanoguanidines is compared with that of the corresponding cyclic urea analogues. The differences in binding constants of the two series of compounds have been rationalized using high-resolution X-ray structure information.  相似文献   

11.
The structural basis of ligand specificity in human immunodeficiency virus (HIV) protease has been investigated by determining the crystal structures of three chimeric HIV proteases complexed with SB203386, a tripeptide analogue inhibitor. The chimeras are constructed by substituting amino acid residues in the HIV type 1 (HIV-1) protease sequence with the corresponding residues from HIV type 2 (HIV-2) in the region spanning residues 31-37 and in the active site cavity. SB203386 is a potent inhibitor of HIV-1 protease (Ki = 18 nM) but has a decreased affinity for HIV-2 protease (Ki = 1280 nM). Crystallographic analysis reveals that substitution of residues 31-37 (30's loop) with those of HIV-2 protease renders the chimera similar to HIV-2 protease in both the inhibitor binding affinity and mode of binding (two inhibitor molecules per protease dimer). However, further substitution of active site residues 47 and 82 has a compensatory effect which restores the HIV-1-like inhibitor binding mode (one inhibitor molecule in the center of the protease active site) and partially restores the affinity. Comparison of the three chimeric protease structures with those of HIV-1 and SIV proteases complexed with the same inhibitor reveals structural changes in the flap regions and the 80's loops, as well as changes in the dimensions of the active site cavity. The study provides structural evidence of the role of the 30's loop in conferring inhibitor specificity in HIV proteases.  相似文献   

12.
The active human immunodeficiency virus type 1 (HIV-1) protease has a homodimeric structure, the subunits are connected by an 'interface' beta-sheet formed by the NH2- and COOH-terminal amino acid segments. Short peptides derived from these segments are able to inhibit the protease activity in the range of micromolar IC50 values. We have further improved the inhibitory power of such peptides by computer modelling. The best inhibitor, the palmitoyl-blocked peptide Pam-Thr-Val-Ser-Tyr-Glu-Leu, has an IC50 value of less than 1 microM. Some of the peptides also showed very good inhibition of the HIV-2 protease. The C-terminal segment of the HIV-1 matrix protein, Acetyl-Gln-Val-Ser-Gln-Asn-Tyr, also inhibits HIV-1 protease. Kinetic studies confirmed the 'dissociative' mechanism of inhibition by the peptides. Depending on the peptide structure and ionic strength, both dimerization inhibition and competitive inhibition were observed, as well as synergistic effects between competitive inhibitors and interface peptides.  相似文献   

13.
To investigate the biochemical properties of the protease encoded by the human endogenous retrovirus, K10 (HERV-K), 213 amino acids of the 3'-end of the HERV-K protease (PR) open reading frame were expressed in Escherichia coli. Autocatalytic cleavage of the expressed polypeptide resulted in an 18.2 kDa protein which was shown to be proteolytically active against a fluorogenic peptide used as a substrate for HIV-1 protease. On the basis of sequence homology and molecular modeling, the 106 N-terminal amino acids of HERV-K PR were predicted to comprise a retroviral protease core domain. An 11.6 kDa protein corresponding to this region was expressed and shown to be a fully functional enzyme. The 11.6 kDa domain of HERV-K PR is unusually stable over a wide pH range, exhibits optimal catalytic activity between pH 4.0 and 5.0, and exists as a dimer at pH 7.0 with a Kd of 50 microM. Like HIV-1 PR, the HERV-K PR core domain is activated by high salt concentrations and processes HIV-1 matrix-capsid polyprotein at the authentic HIV-1 PR recognition site. However, both the 18.2 and 11.6 kDa forms of HERV-K PR were highly resistant to a number of clinically useful HIV-1 PR inhibitors, including ritonavir, indinavir, and saquinavir. This raises the possibility that HERV-K PR may complement HIV-1 PR during infection, and could have implications for protease inhibitor therapy and drug resistance.  相似文献   

14.
The proteases expressed by the HIV-1 and HIV-2 viruses process the polyproteins encoded by the viral genomes into the mature proteins required for virion replication and assembly. Eight analogs of haloperidol have been synthesized that cause time-dependent inactivation of the HIV-1 protease and, in six cases, HIV-2 protease. The IC50 values for the analogues are comparable to that of haloperidol itself. Enzyme inactivation is due to the presence of an epoxide in two of the analogues and carbonyl-conjugated double or triple bonds in the others. Irreversible inactivation is confirmed by the failure to recover activity when one of the inhibitors is removed from the medium. At pH 8.0, the agents inactivate the HIV-1 protease 4-80 times more rapidly than the HIV-2 protease. Faster inactivation of the HIV-1 protease is consistent with alkylation of cysteine residues because the HIV-1 protease has four such residues whereas the HIV-2 protease has none. Inactivation of the HIV-2 protease requires modification of non-cysteine residues. The similarities in the rates of inactivation of the HIV-2 protease by six agents that have intrinsically different reactivities toward nucleophiles suggest that the rate-limiting step in the inactivation process is not the alkylation reaction itself. At least five of the agents inhibit polyprotein processing in an ex vivo cell assay system, but they are also toxic to the cells.  相似文献   

15.
Protease inhibitors are currently the most effective antiviral agents against human immunodeficiency virus type 1 (HIV-1). In this study we determined the effect of four HIV-1 protease inhibitors on human T cell leukemia virus type 1 (HTLV-I). Rhesus monkey cells infected with HTLV-I were treated with different concentrations of indinavir, saquinavir, ritonavir, or nelfinavir. The effect of these inhibitors was monitored through their effect on the processing efficiency of the viral Gag protein in cells, the natural substrate for the viral protease. These inhibitors failed to block processing of HTLV-I Gag. To confirm these findings, human cells were cotransfected with plasmids encoding infectious copies of HIV-1 and HTLV-I, and the cells were subsequently treated with these same HIV-1 protease inhibitors. At concentrations between 5 and 50 times the IC50 for inhibition of HIV-1 replication, inhibition of HIV-1 Gag cleavage was apparent. In contrast, no effect on HTLV-I Gag processing was seen. At higher concentrations, HIV-1 Gag processing was essentially completely inhibited whereas HTLV-I Gag cleavage was still unaffected. Thus, these inhibitors are not effective inhibitors of HTLV-I Gag processing. Sequence alignments of the HIV-1 and HTLV-I viral proteases and processing sites suggest that the active site of the HTLV-I protease may have subtle differences in substrate recognition compared with the HIV-1 protease.  相似文献   

16.
Toward establishing the general efficacy of using trisubstituted cyclopropanes as peptide mimics to stabilize extended peptide structures, the cyclopropanes 20a-d were incorporated as replacements into 9-13, which are analogues of the known HIV-1 protease inhibitors 14 and 15. The syntheses of 20a-d commenced with the Rh2[5(S)-MEPY]4-catalyzed cyclization of the allylic diazoesters 16a-d to give the cyclopropyl lactones 17a-d in high enantiomeric excess. Opening of the lactone moiety using the Weinreb protocol and straightforward refunctionalization of the intermediate amides 18a-d gave 20a-d. A similar sequence of reactions was used to prepare the N-methyl-2-pyridyl analogue 28. Coupling of 20a-d and 28 with the known diamino diol 22 delivered 9-13. Pseudopeptides 9-12 were found to be competitive inhibitors of wild-type HIV-1 protease in biological assays having Kis of 0.31-0.35 nM for 9, 0.16-0.21 nM for 10, 0.47 nM for 11, and 0.17 nM for 12; these inhibitors were thus approximately equipotent to the known inhibitor 14(IC50 = 0.22 nM) from which they were derived. On the other hand 13 (Ki = 80 nM) was a weaker inhibitor than its analogue 15 (Ki = 0.11 nM). The solution structures of 9 and 10 were analyzed by NMR spectroscopy and simulated annealing procedures that included restraints derived from homo- and heteronuclear coupling constants and NOEs; because of the molecular symmetry of9 and 10, a special protocol to treat the NOE data was used. The final structure was checked by restrained and free molecular dynamic calculations using an explicit DMSO solvent box. The preferred solution conformations of 9 and 10 are extended structures that closely resemble the three-dimensional structure of 10 bound to HIV-1 protease as determined by X-ray crystallographic analysis of the complex. This work convincingly demonstrates that extended structures of peptides may be stabilized by the presence of substituted cyclopropanes that serve as peptide replacements. Moreover, the linear structure enforced in solution by the two cyclopropane rings in the pseudopeptides 9-12 appears to correspond closely to the biologically active conformation of the more flexible inhibitors 14 and 15. The present work, which is a combination of medicinal, structural, and quantum chemistry, thus clearly establishes that cyclopropanes may be used as structural constraints to reduce the flexibility of linear pseudopeptides and to help enforce the biologically active conformation of such ligands in solution.  相似文献   

17.
The design, synthesis, and crystallographic analysis of protein-inhibitor complexes is described for a novel series of nonpeptidic HIV protease (HIV Pr)inhibitors. Beginning with a cocrystal structure of a Phe-Pro peptidomimetic bound to the HIV Pr, design was initiated that resulted in the substituted 2-butanol compound 8 as the lead compound (Ki = 24.5 microM, racemic mixture). Modifications on the initial compound were then made on the basis of its cocrystal structure with HIV Pr and inhibition data, resulting in compounds with enhanced potency against the enzyme (compound 18, Ki = 0.48 microM). These inhibitors were found to bind to the enzyme essentially as predicted on the basis of the original design hypothesis. Stereospecific synthesis of individual enantiomers confirmed the prediction of a binding preference for the S alcohol stereochemistry. Modest antiviral activity was demonstrated for several of the more potent HIV Pr inhibitors in a HIV-1 infected CEM-SS cell line.  相似文献   

18.
Human seminal plasma contains two acid-stable proteinase inhibitors, HUSI-II (Mr approximately 6500) and HUSI-I, (Mr approximately 11 000) with different inhibition specificities. The inhibitory activity of HUSI-II is strongly limited to trypsin and acrosin; both enzyme-inhibitor complexes are very stable (e.g. bovine trypsin-HUSI-II complex: Ki = 1 x 10(-10)M; human acrosin-HUSI-II complex: Ki = 2.7 x 10(-10)M). The inhibitor from human seminal plasma HUSI-II may therefore be seen as the natural antagonist of the sperm protease acrosin. In addition to pancreatic trypsin and alpha-chymotrypsin, HUSI-I forms strong complexes with neutral proteases of the lysosome-like granules from human granulocytes, for example, the elastase (Ki = 2.5 x 10(-9)M) and cathepsin G, the chymotrypsin like protease (Ki = 7 x 10(-8)M).  相似文献   

19.
OBJECTIVE: To define genotypic and phenotypic resistance patterns following prolonged therapy with the protease inhibitor ritonavir (ABT-538). DESIGN: Seven HIV-1-infected patients, all but one previously treated with dideoxynucleoside analogues (zidovudine, didanosine, zalcitabine), were treated for 1 year with ritonavir. METHODS: Direct solid-phase sequencing of the protease gene starting from plasma derived viral RNA followed by comparison to phenotypic drug resistance data. RESULTS: The most frequent amino-acid substitutions occurring upon administration of the protease inhibitor were V82A/F (substrate binding site), I54V (flap region), A71V and L10I. Additional mutations found in more than one patient were I15V, M36I, I84V and I93L. Mutation L63P was found both in pre- and post-ritonavir samples. Phenotypic drug resistance assays confirmed resistance to ritonavir in post-treatment samples (approximately 170-fold) and showed cross-resistance to indinavir (approximately 30-fold) and partially to saquinavir (approximately fivefold). At 1 year of treatment, one patient without known resistance-associated mutations in the protease gene still showed a substantial rise in CD4 cell count accompanied by a more than 2.4 log decrease in RNA viral load. However, at week 78, mutations R8Q, E34K, R57K, L63P and I84V were detected and the treatment benefit was partially lost. CONCLUSIONS: Long-term treatment with ritonavir is associated with the emergence of multiple mutations in the HIV-1 protease gene. The mutations L10I, I54V, L63P, A71V, V82A/F and I84V correspond to known drug-resistance mutations for ritonavir and other protease inhibitors. Phenotypic resistance to ritonavir was detected in a majority of ritonavir-treated patients at 1 year of treatment. In addition, long-term ritonavir treatment selects for cross-resistance to the protease inhibitors indinavir and saquinavir. This argues against sequential therapy with several protease inhibitors. Delayed resistance in one patient was accompanied with a prolonged increase in CD4 cell count and decrease in viral load suggesting a temporary benefit of treatment.  相似文献   

20.
Rigid body rotation of five domains and movements within their interfacial joints provide a rational context for understanding why HIV protease mutations that arise in drug resistant strains are often spatially removed from the drug or substrate binding sites. Domain motions associated with substrate binding in the retroviral HIV-1 and SIV proteases are identified and characterized. These motions are in addition to closure of the flaps and result from rotations of approximately 6-7 degrees at primarily hydrophobic interfaces. A crystal structure of unliganded SIV protease (incorporating the point mutation Ser 4 His to stabilize the protease against autolysis) was determined to 2.0 A resolution in a new space group, P3221. The structure is in the most "open" conformation of any retroviral protease so far examined, with six residues of the flaps disordered. Comparison of this and unliganded HIV structures, with their respective liganded structures by difference distance matrixes identifies five domains of the protease dimer that move as rigid bodies against one another: one terminal domain encompassing the N- and C-terminal beta sheet of the dimer, two core domains containing the catalytic aspartic acids, and two flap domains. The two core domains rotate toward each other on substrate binding, reshaping the binding pocket. We therefore show that, for enzymes, mutations at interdomain interfaces that favor the unliganded form of the target active site will increase the off-rate of the inhibitor, allowing the substrate greater access for catalysis. This offers a mechanism of resistance to competitive inhibitors, especially when the forward enzymatic reaction rate exceeds the rate of substrate dissociation.  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