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.
As a continuation of our efforts to discover and develop the apoptosis inducing 1-benzoyl-3-cyanopyrrolo[1,2-a]quinolines as potential anticancer agents, we explored substitutions at the 4-, 5-, 6-, 7- and 8-positions of pyrrolo[1,2-a]quinoline. SAR studies showed that substitution at the 6-position by a small group such as Cl resulted in potent compounds. Substitutions at the 5- and 8-positions were tolerated while substitutions at the 4- and 7-position led to inactive compounds. Several compounds, including 2c, 3a, 3b and 3f, were found to be highly active against human breast cancer cells T47D with EC50 values of 0.053–0.080 μM, but much less active against human colon cancer cells HCT116 and hepatocellular carcinoma cancer cells SNU398 in the caspase activation assay. Compound 3f also was found to be highly active with a GI50 value of 0.018 μM against T47D cells in a growth inhibition assay.  相似文献   

2.
A range of 3,6-di(hetero)arylimidazo[1,2-a]pyrazine ATP-competitive inhibitors of CHK1 were developed by scaffold hopping from a weakly active screening hit. Efficient synthetic routes for parallel synthesis were developed to prepare analogues with improved potency and ligand efficiency against CHK1. Kinase profiling showed that the imidazo[1,2-a]pyrazines could inhibit other kinases, including CHK2 and ABL, with equivalent or better potency depending on the pendant substitution. These 3,6-di(hetero)aryl imidazo[1,2-a]pyrazines appear to represent a general kinase inhibitor scaffold.  相似文献   

3.
The environmental distribution of Dehalococcoides group organisms and their association with chloroethene-contaminated sites were examined. Samples from 24 chloroethene-dechlorinating sites scattered throughout North America and Europe were tested for the presence of members of the Dehalococcoides group by using a PCR assay developed to detect Dehalococcoides 16S rRNA gene (rDNA) sequences. Sequences identified by sequence analysis as sequences of members of the Dehalococcoides group were detected at 21 sites. Full dechlorination of chloroethenes to ethene occurred at these sites. Dehalococcoides sequences were not detected in samples from three sites at which partial dechlorination of chloroethenes occurred, where dechlorination appeared to stop at 1,2-cis-dichloroethene. Phylogenetic analysis of the 16S rDNA amplicons confirmed that Dehalococcoides sequences formed a unique 16S rDNA group. These 16S rDNA sequences were divided into three subgroups based on specific base substitution patterns in variable regions 2 and 6 of the Dehalococcoides 16S rDNA sequence. Analyses also demonstrated that specific base substitution patterns were signature patterns. The specific base substitutions distinguished the three sequence subgroups phylogenetically. These results demonstrated that members of the Dehalococcoides group are widely distributed in nature and can be found in a variety of geological formations and in different climatic zones. Furthermore, the association of these organisms with full dechlorination of chloroethenes suggests that they are promising candidates for engineered bioremediation and may be important contributors to natural attenuation of chloroethenes.  相似文献   

4.
L-Homophenylalanine (L-HPhe) is the precursor of 2-phenylethylglucosinolate, a secondary metabolite present in some Brassica and related species. A key step in its biosynthesis is the oxidative decarboxylation of L-HPhe to its aldoxime. The enzyme catalysing this reaction has been shown to be a NADPH- and O2-dependent microsomal flavoprotein (L-HPhe FP; EC unclassified). Inhibition studies using Phe homologs and HPhe analogs (α-amino-, α-carboxyl- and ring-substituted), and specific amino acid modifications, were carried out to determine the possible active site structure and catalytic mechanism of L-HPhe FP. Activity with L-HPhe was inhibited by the two higher homologs, but not by L-Phe. Methylation of the substrate α-amino group, or replacement of the α-carboxyl group with a phosphonic acid group, significantly reduced the inhibition. Ring substitutions had varying effects: single methyl substitutions had only minor effects on binding to the active site, whereas di- or tri-methyl, methoxy or halide substitutions significantly reduced inhibition. Simple amines had no significant effect on L-HPhe FP activity. Binding to the active site of the enzyme appears to require a minimum chain length, plus an aromatic ring at one end of the molecule and unmodified α-amino acid moiety at the other. Chemical modification of amino acids on the protein implied there was no requirement for thiol groups (-SH), Ser/Thr hydroxyl groups, or L-Arg in the active site of L-HPhe FP. However, there was evidence for the presence of essential His and Tyr residues, and the involvement of Glu or Asp residues at or near the active site.  相似文献   

5.
Nitrobenzene 1,2-dioxygenase from Comamonas sp. strain JS765 catalyzes the initial reaction in nitrobenzene degradation, forming catechol and nitrite. The enzyme also oxidizes the aromatic rings of mono- and dinitrotoluenes at the nitro-substituted carbon, but the basis for this specificity is not understood. In this study, site-directed mutagenesis was used to modify the active site of nitrobenzene dioxygenase, and the contribution of specific residues in controlling substrate specificity and enzyme performance was evaluated. The activities of six mutant enzymes indicated that the residues at positions 258, 293, and 350 in the α subunit are important for determining regiospecificity with nitroarene substrates and enantiospecificity with naphthalene. The results provide an explanation for the characteristic specificity with nitroarene substrates. Based on the structure of nitrobenzene dioxygenase, substitution of valine for the asparagine at position 258 should eliminate a hydrogen bond between the substrate nitro group and the amino group of asparagine. Up to 99% of the mononitrotoluene oxidation products formed by the N258V mutant were nitrobenzyl alcohols rather than catechols, supporting the importance of this hydrogen bond in positioning substrates in the active site for ring oxidation. Similar results were obtained with an I350F mutant, where the formation of the hydrogen bond appeared to be prevented by steric interference. The specificity of enzymes with substitutions at position 293 varied depending on the residue present. Compared to the wild type, the F293Q mutant was 2.5 times faster at oxidizing 2,6-dinitrotoluene while retaining a similar Km for the substrate based on product formation rates and whole-cell kinetics.  相似文献   

6.
Two-component regulatory systems, in which phosphorylation controls the activity of a response regulator protein, provide signal transduction in bacteria. For example, the phosphorylated CheY response regulator (CheYp) controls swimming behavior. In Escherichia coli, the chemotaxis phosphatase CheZ stimulates the dephosphorylation of CheYp. CheYp apparently binds first to the C terminus of CheZ and then binds to the active site where dephosphorylation occurs. The phosphatase activity of the CheZ2 dimer exhibits a positively cooperative dependence on CheYp concentration, apparently because the binding of the first CheYp to CheZ2 is inhibited compared to the binding of the second CheYp. Thus, CheZ phosphatase activity is reduced at low CheYp concentrations. The CheZ21IT gain-of-function substitution, located far from either the CheZ active site or C-terminal CheY binding site, enhances CheYp binding and abolishes cooperativity. To further explore mechanisms regulating CheZ activity, we isolated 10 intragenic suppressor mutations of cheZ21IT that restored chemotaxis. The suppressor substitutions were located along the central portion of CheZ and were not allele specific. Five suppressor mutants tested biochemically diminished the binding of CheYp and/or the catalysis of dephosphorylation, even when the suppressor substitutions were distant from the active site. One suppressor mutant also restored cooperativity to CheZ21IT. Consideration of results from this and previous studies suggests that the binding of CheYp to the CheZ active site (not to the C terminus) is rate limiting and leads to cooperative phosphatase activity. Furthermore, amino acid substitutions distant from the active site can affect CheZ catalytic activity and CheYp binding, perhaps via the propagation of structural or dynamic perturbations through a helical bundle.  相似文献   

7.
The two transesterification reactions catalyzed by self-splicing group II introns take place in either two active sites or two conformations of a single active site involving rearrangements of the positions of the reacting groups. We have investigated the effects on the rates of the chemical steps of the two reactions due to sulfur substitution of nonbridging oxygens at both the 5' and 3' splice sites as well as the deoxyribose substitution of the ribose 2' hydroxyl group at the 5' splice site. The data suggest that the two active sites differ in their interactions with several of these groups. Specifically, sulfur substitution of the pro-Sp nonbridging oxygen at the 5' splice site reduces the chemical rate of the step one branching reaction by at least 250-fold, whereas substitution of the pro-Sp oxygen at the 3' splice site has only a 4.5-fold effect on the chemical rate of step two. Previous work demonstrated that the Rp phosphorothioate substitutions at both the 5' and 3' splice sites reduced the rate of both steps of splicing to an undetectable level. These results suggest that either two distinct active sites catalyze the two steps or that more significant alterations must be made in a single bifunctional active site to accommodate the two different reactions.  相似文献   

8.
The functional roles of the two nucleotide binding folds, NBF1 and NBF2, in the activation of the cystic fibrosis transmembrane conductance regulator (CFTR) were investigated by measuring the rates of activation and deactivation of CFTR Cl- conductance in Xenopus oocytes. Activation of wild-type CFTR in response to application of forskolin and 3-isobutyl-1-methylxanthine (IBMX) was described by a single exponential. Deactivation after washout of the cocktail consisted of two phases: an initial slow phase, described by a latency, and an exponential decline. Rate analysis of CFTR variants bearing analogous mutations in NBF1 and NBF2 permitted us to characterize amino acid substitutions according to their effects on the accessibility and stability of the active state. Access to the active state was very sensitive to substitutions for the invariant glycine (G551) in NBF1, where mutations to alanine (A), serine (S), or aspartic acid (D) reduced the apparent on rate by more than tenfold. The analogous substitutions in NBF2 (G1349) also reduced the on rate, by twofold to 10-fold, but substantially destabilized the active state as well, as judged by increased deactivation rates. In the putative ATP-binding pocket of either NBF, substitution of alanine, glutamine (Q), or arginine (R) for the invariant lysine (K464 or K1250) reduced the on rate similarly, by two- to fourfold. In contrast, these analogous substitutions produced opposite effects on the deactivation rate. NBF1 mutations destabilized the active state, whereas the analogous substitutions in NBF2 stabilized the active state such that activation was prolonged compared with that seen with wild-type CFTR. Substitution of asparagine (N) for a highly conserved aspartic acid (D572) in the ATP-binding pocket of NBF1 dramatically slowed the on rate and destabilized the active state. In contrast, the analogous substitution in NBF2 (D1370N) did not appreciably affect the on rate and markedly stabilized the active state. These results are consistent with a hypothesis for CFTR activation that invokes the binding and hydrolysis of ATP at NBF1 as a crucial step in activation, while at NBF2, ATP binding enhances access to the active state, but the rate of ATP hydrolysis controls the duration of the active state. The relatively slow time courses for activation and deactivation suggest that slow processes modulate ATP-dependent gating.  相似文献   

9.
As a part of ongoing studies in developing novel anticancer agents, a series of modified 2,4-diaryl-5H-indeno[1,2-b]pyridines were designed, and synthesized by introducing hydroxyl and chlorine moieties. They were evaluated for topoisomerase inhibitory activity and cytotoxicity against HCT15, T47D, and HeLa cancer cell lines. This modification allowed us to demonstrate structure–activity relationship (SAR) study with respect to the non-substituted 2,4-diaryl-5H-indeno[1,2-b]pyridines. Compounds (2, 3, 4, 5, 8, and 9) with meta or para hydroxyl group on 2 or 4-phenyl ring have enhanced topo I and II inhibitory activity and cytotoxicity. However, additional substitution of chlorine group on furyl or thienyl ring (11, 12, 14, 16–18) generally reduced topo I and II inhibitory activity but improved cytotoxicity. The observation of cytotoxic properties and SAR study according to the position of hydroxyl and chlorine group will provide valuable insight for further study of development of novel anticancer agents with related scaffolds.  相似文献   

10.
Critical to human innate immunity against African trypanosomes is a minor subclass of human high-density lipoproteins, termed Trypanosome Lytic Factor-1 (TLF-1). This primate-specific molecule binds to a haptoglobin-hemoglobin receptor (HpHbR) on the surface of susceptible trypanosomes, initiating a lytic pathway. Group 1 Trypanosoma brucei gambiense causes human African Trypanosomiasis (HAT), escaping TLF-1 killing due to reduced uptake. Previously, we found that group 1 T. b. gambiense HpHbR (TbgHpHbR) mRNA levels were greatly reduced and the gene contained substitutions within the open reading frame. Here we show that a single, highly conserved amino acid in the TbgHpHbR ablates high affinity TLF-1 binding and subsequent endocytosis, thus evading TLF-1 killing. In addition, we show that over-expression of TbgHpHbR failed to rescue TLF-1 susceptibility. These findings suggest that the single substitution present in the TbgHpHbR directly contributes to the reduced uptake and resistance to TLF-1 seen in these important human pathogens.  相似文献   

11.
We compared two haploid genotypes of one Ciona savignyi individual and identified codons at which these genotypes differ by two nonsynonymous substitutions. Using the C. intestinalis genome as an outgroup, we showed that both substitutions tend to occur in the same genotype. Only in 53 (34.4%) of 154 codons, one substitution occurred in each of the two genotypes, although 77 (50%) of such codons are to be expected if substitutions were independent. We considered two feasible evolutionary causes for the observed pattern: substitutions driven by positive selection and compensatory substitutions, as well as several potential biases. However, none of these explanations is fully compelling, and data on multiple genotypes of C. savignyi would help to elucidate the causes of this pattern.  相似文献   

12.
Type II R67 dihydrofolate reductase (DHFR) is a bacterial plasmid-encoded enzyme that is intrinsically resistant to the widely-administered antibiotic trimethoprim. R67 DHFR is genetically and structurally unrelated to E. coli chromosomal DHFR and has an unusual architecture, in that four identical protomers form a single symmetrical active site tunnel that allows only one substrate binding/catalytic event at any given time. As a result, substitution of an active-site residue has as many as four distinct consequences on catalysis, constituting an atypical model of enzyme evolution. Although we previously demonstrated that no single residue of the native active site is indispensable for function, library selection here revealed a strong bias toward maintenance of two native protomers per mutated tetramer. A variety of such “half-native” tetramers were shown to procure native-like catalytic activity, with similar KM values but kcat values 5- to 33-fold lower, illustrating a high tolerance for active-site substitutions. The selected variants showed a reduced thermal stability (Tm ∼12°C lower), which appears to result from looser association of the protomers, but generally showed a marked increase in resilience to heat denaturation, recovering activity to a significantly greater extent than the variant with no active-site substitutions. Our results suggest that the presence of two native protomers in the R67 DHFR tetramer is sufficient to provide native-like catalytic rate and thus ensure cellular proliferation.  相似文献   

13.
M1 family metallo-aminopeptidases fulfill a wide range of critical and in some cases medically relevant roles in humans and human pathogens. The specificity of M1-aminopeptidases is dominated by the interaction of the well defined S1 subsite with the side chain of the first (P1) residue of the substrate and can vary widely. Extensive natural variation occurs at one of the residues that contributes to formation of the cylindrical S1 subsite. We investigated whether this natural variation contributes to diversity in S1 subsite specificity. Effects of 11 substitutions of the S1 subsite residue valine 459 in the Plasmodium falciparum aminopeptidase PfA-M1 and of three substitutions of the homologous residue methionine 260 in Escherichia coli aminopeptidase N were characterized. Many of these substitutions altered steady-state kinetic parameters for dipeptide hydrolysis and remodeled S1 subsite specificity. The most dramatic change in specificity resulted from substitution with proline, which collapsed S1 subsite specificity such that only substrates with P1-Arg, -Lys, or -Met were appreciably hydrolyzed. The structure of PfA-M1 V459P revealed that the proline substitution induced a local conformational change in the polypeptide backbone that resulted in a narrowed S1 subsite. The restricted specificity and active site backbone conformation of PfA-M1 V459P mirrored those of endoplasmic reticulum aminopeptidase 2, a human enzyme with proline in the variable S1 subsite position. Our results provide compelling evidence that changes in the variable residue in the S1 subsite of M1-aminopeptidases have facilitated the evolution of new specificities and ultimately novel functions for this important class of enzymes.  相似文献   

14.
Tropomyosin polymerizes along actin filaments and together with troponin regulates muscle contraction in a Ca-dependent manner. Actin-binding periods are homologous residues, which repeat along tropomyosin sequence, form tropomyosin-actin interface and determine regulatory functions. To learn how period 3 is involved in tropomyosin functions we examined effects of two mutations in Tpm1.1, I92T and V95A, which have been linked to dilated and hypertrophic cardiomyopathies characterized respectively by hyper- and hypocontractile phenotypes. In this work the functional consequences of both mutations were studied in vitro by using actin thin filaments reconstituted in the presence of mutant Tpm1.1 homodimers carrying the substitutions in both tropomyosin chains, Tpm1.1 heterodimers with substitution only in one Tpm1.1 chain, and Tpm1.1/Tpm2.2 heterodimers with substitution in Tpm1.1 chain and wild type Tpm2.2 in the second chain. The presence of the substitution I92T decreased the tropomyosin affinity for actin, abolished Ca2+-dependent activation of the actomyosin ATPase, decreased the sensitivity of the tropomyosin-troponin complex to subsaturating Ca2+ concentrations and reduced the cooperativity of the myosin-induced transition of the thin filament to a fully active state. The substitution V95A had opposite effects: increased actin affinity, increased the actomyosin ATPase activity above the level observed for wild type Tpm and increased cooperativity of myosin-induced activation of the thin filaments reconstructed with homo- and heterodimers of tropomyosin. Substitutions I92T and V95A were dominant, but the formation of heterodimers modified the effects observed for homodimers.  相似文献   

15.
Prion protein (PrP) contains two N-linked glycosylation sites. It is unknown which amino acid substitution contributes most efficiently to the abolishment of N-linked glycosylations. To define the influence of amino acid substitution at the N-linked glycosylation sites on the conversion efficiency of mouse PrP, we tested each of all 19 amino acid substitutions at either one of the N-linked glycosylation sites (codon 180, 182, 196 or 198). The conversion efficiency of the mutagenized PrP was highly dependent on the newly introduced amino acid itself regardless of the absence of N-linked glycosylation in scrapie-infected mouse neuroblastoma cells. The majority of mutant PrP with substitutions at the Asn residues of the N-linked glycosylation sites were conversion-competent, whereas most mutant PrP with substitutions at the Thr residues were conversion-incompetent. These findings emphasize that the Asn residues of the N-linked glycosylation sites are replaceable to abolish N-linked glycosylations without directly affecting the protein function.  相似文献   

16.
A total of 790 Drosophila melanogaster genes that are alternatively spliced in a coding region and have orthologs in Drosophila pseudoobscura were studied. It proved that nucleotide substitutions are accumulated in alternative coding regions more rapidly than in constitutive coding regions. Moreover, the evolutionary patterns of alternative regions differing in insertion-deletion mechanisms (use of alternative promoters, splicing sites, or polyadenylation sites) differ significantly. The synonymous substitution rate in coding regions of genes varies more strongly than the nonsynonymous substitution rate. The patterns of substitutions in different classes of alternative regions of Drosophila melanogaster and mammals differ considerably.  相似文献   

17.
CheY serves as a structural prototype for the response regulator proteins of two-component regulatory systems. Functional roles have previously been defined for four of the five highly conserved residues that form the response regulator active site, the exception being the hydroxy amino acid which corresponds to Thr87 in CheY. To investigate the contribution of Thr87 to signaling, we characterized, genetically and biochemically, several cheY mutants with amino acid substitutions at this position. The hydroxyl group appears to be necessary for effective chemotaxis, as a Thr→Ser substitution was the only one of six tested which retained a Che+ swarm phenotype. Although nonchemotactic, cheY mutants with amino acid substitutions T87A and T87C could generate clockwise flagellar rotation either in the absence of CheZ, a protein that stimulates dephosphorylation of CheY, or when paired with a second site-activating mutation, Asp13→Lys, demonstrating that a hydroxy amino acid at position 87 is not essential for activation of the flagellar switch. All purified mutant proteins examined phosphorylated efficiently from the CheA kinase in vitro but were impaired in autodephosphorylation. Thus, the mutant CheY proteins are phosphorylated to a greater degree than wild-type CheY yet support less clockwise flagellar rotation. The data imply that Thr87 is important for generating and/or stabilizing the phosphorylation-induced conformational change in CheY. Furthermore, the various position 87 substitutions differentially affected several properties of the mutant proteins. The chemotaxis and autodephosphorylation defects were tightly linked, suggesting common structural elements, whereas the effects on self-catalyzed and CheZ-mediated dephosphorylation of CheY were uncorrelated, suggesting different structural requirements for the two dephosphorylation reactions.  相似文献   

18.
The first symmetry by base substitutions of degeneracy in the genetic code was described by Rumer (1966) and the other symmetries were identified later by Jestin (2006) and Jestin and Soulé (2007). Here, a rationale accounting for these symmetries is reported. The number of non-synonymous substitutions over the replicated coding sequence is written as a function of the substitution matrix, whose elements are the number of substitutions from any codon to any other codon. The p-adic distance used as a similarity measure and applied to this matrix is shown to be biologically relevant. The rationale indicates that symmetries by base substitutions of degeneracy in the genetic code are symmetries of the measures of the number of non-synonymous substitutions for sets of synonymous codons.  相似文献   

19.
The properties of oxidation of dichloroethene (DCE) and trichloroethylene (TCE) by three mutant strains of Pseudomonas butanovora containing single amino acid substitutions in the α-subunit of butane monooxygenase hydroxylase (BMOH-α) were compared to the properties of the wild-type strain (Rev WT). The rates of oxidation of three chloroethenes (CEs) were reduced in mutant strain G113N and corresponded with a lower maximum rate of butane oxidation. The rate of TCE degradation was reduced by one-half in mutant strain L279F, whereas the rates of DCE oxidation were the same as those in Rev WT. Evidence was obtained that the composition of products of CE oxidation differed between Rev WT and some of the mutant strains. For example, while Rev WT released nearly all available chlorine stoichiometrically during CE oxidation, strain F321Y released about 40% of the chlorine during 1,2-cis-DCE and TCE oxidation, and strain G113N released between 14 and 25% of the available chlorine during oxidation of DCE and 56% of the available chlorine during oxidation of TCE. Whereas Rev WT, strain L279F, and strain F321Y formed stoichiometric amounts of 1,2-cis-DCE epoxide during oxidation of 1,2-cis-DCE, only about 50% of the 1,2-cis-DCE oxidized by strain G113N was detected as the epoxide. Evidence was obtained that 1,2-cis-DCE epoxide was a substrate for butane monooxygenase (BMO) that was oxidized after the parent compound was consumed. Yet all of the mutant strains released less than 40% of the available 1,2-cis-DCE chlorine, suggesting that they have altered activity towards the epoxide. In addition, strain G113N was unable to degrade the epoxide. TCE epoxide was detected during exposure of Rev WT and strain F321Y to TCE but was not detected with strains L279F and G113N. Lactate-dependent O2 uptake rates were differentially affected by DCE degradation in the mutant strains, providing evidence that some products released by the altered BMOs reduced the impact of CE on cellular toxicity. The use of CEs as substrates in combination with P. butanovora BMOH-α mutants might allow insights into the catalytic mechanism of BMO to be obtained.  相似文献   

20.
The effects of ion substitutions on the Cl- secretion rate and tissue conductance of isolated short-circuited opercular epithelia from sea-water-adapted Fundulus heteroclitus were investigated. Serosal Na+ substitution had the same effect on the Cl- secretion rate that serosal Cl- substitution had on the active component of the Cl- efflux. This similarity indicated a 1:1 Na-Cl requirement for active Cl- secretion across this epithelium, which supports the proposal of a coupled NaCl uptake mechanism at the serosal membrane of Cl- secretory epithelia. Mucosal Na+ and Cl- substitutions appeared to inhibit completely the active Cl- secretory flux. The reductions in the tissue conductance with mucosal ion substitutions suggested that this effect can be attributed to a blocking of the apical membrane Cl- conductance. These mucosal ion effects suggested a possible direct regulatory influence of the external salinity on the Cl- secretion rate and tissue conductance, which provide alternative explanations for observations with the teleost gill epithelium.  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