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.
Ravi M  Zask A  Rush TS 《Biochemistry》2005,44(48):15871-15879
A binding mode of HTI-286, a synthetic analogue of the peptidic antimitotic agent hemiasterlin, to tubulin is proposed. The binding mode was derived from iterative docking experiments directed at regions of the tubulin interdimer interface that are believed to be consistent with all current experimental data regarding the HTI-286/tubulin interaction. These data include (1) competitive inhibition of the tubulin binding of the Vinca alkaloids and other antimitotic agents, (2) proximity to stretches of amino acid residues identified in two separate photoaffinity-labeling experiments, (3) structure-activity relationships for HTI-286 and its analogues, (4) saturation transfer difference nuclear magnetic resonance (NMR) experiments, and (5) NMR transfer nuclear Overhauser effect spectroscopy (NOESY) experiments that potentially identify the bioactive conformation. The predicted binding mode thus affords a means to understand the mode of action of hemiasterlin, HTI-286, and other closely related molecules.  相似文献   

2.
HTI-286 is a synthetic analogue of the natural product hemiasterlin and is a potent antimitotic agent. HTI-286 inhibits the proliferation of tumor cells during mitosis. The observed antimitotic activity is due to the binding of HTI-286 to tubulin. This report details the effects of HTI-286 on soluble tubulin and preassembled microtubules. HTI-286 binds tubulin monomer and oligomerizes it to an 18.5 S species corresponding to a discrete ring structure consisting of about 13 tubulin units as determined by sedimentation equilibrium analyses. The rate of formation of the oligomers is dependent on the concentration of HTI-286 and the time of incubation. Tubulin oligomers, specifically the 18.5 S species, form slowly. The interactions of HTI-286 with tubulin were studied by isothermal titration calorimetry. HTI-286 binds tubulin rapidly, and the initial association of HTI-286 with tubulin is enthalpically driven with a DeltaH value of -14 kcal/mol at 25 degrees C and a dissociation constant of ca. 100 nM. However, the accompanying tubulin oligomerization event does not produce measurable heats at 25 degrees C. The dissociation constant estimated from the changes in the intrinsic fluorescence of tubulin was found to be consistent with the calorimetric results. Both HTI-286 and hemiasterlin bind tubulin with nearly equal potency. However, the stability of the tubulin oligomers is not identical under size-exclusion column chromatographic conditions. The tubulin oligomers formed in the presence of HTI-286 dissociate on the column, while the corresponding oligomers formed in the presence of hemiasterlin are stable. Tubulin undergoes a change in the secondary structure in the presence of HTI-286, which is evidenced by changes in the circular dichroic absorption spectrum of tubulin. In contrast to the microtubule-stabilizing effects of paclitaxel, both HTI-286 and hemiasterlin depolymerize preassembled microtubules at micromolar concentrations.  相似文献   

3.
A synthetic analogue of the tripeptide hemiasterlin, designated HTI-286, depolymerizes microtubules, is a poor substrate for P-glycoprotein, and inhibits the growth of paclitaxel-resistant tumors in xenograft models. Two radiolabeled photoaffinity analogues of HTI-286, designated 4-benzoyl-N,beta,beta-trimethyl-l-phenylalanyl-N(1)-[(1S,2E)-3-carboxy-1-isopropylbut-2-enyl]-N(1),3-dimethyl-l-valinamide (probe 1) and N,beta,beta-trimethyl-l-phenylalanyl-4-benzoyl-N-[(1S,2E)-3-carboxy-1-isopropyl-2-butenyl]-N,beta,beta-trimethyl-l-phenylalaninamide (probe 2), were made to help identify HTI-286 binding sites in tubulin. HTI-286, probe 1, and probe 2 had similar affinities for purified tubulin [apparent K(D(app)) = 0.2-1.1 microM], inhibited polymerization of purified tubulin approximately 80%, and were potent inhibitors of cell growth (IC(50) = 1.0-22 nM). Both radiolabeled probes labeled exclusively alpha-tubulin. Labeling by [(3)H]probe 1 was inhibited by probe 1, HTI-286, vinblastine, or dolastatin 10 (another peptide antimitotic agent that depolymerizes microtubules) but was either unaffected or enhanced (at certain temperatures) by colchicine or paclitaxel. [(3)H]Probe 1 also labeled exclusively tubulin in cytosolic extracts of whole cells. The major, if not exclusive, contact site for probe 1 was mapped to residues 314-339 of alpha-tubulin and corresponds to the sheet 8 and helix 10 region. This region is known to (1) have longitudinal interactions with beta-tubulin across the interdimer interface, (2) have lateral interactions with adjacent protofilaments, and (3) contact the N-terminal region of stathmin, a protein that induces depolymerization of tubulin. Binding of probe 1 to this region may alter the conformation of tubulin outside the labeling domain, since enzymatic removal of the C-terminus of only alpha-tubulin by subtilisin after, but not before, photolabeling is blocked by probe 1. These results suggest that hemiasterlin is in close contact with alpha-tubulin and may span the interdimer interface so that it contacts the vinblastine- and dolastatin 10-binding sites believed to be in beta-tubulin. In addition, we speculate that antimitotic peptides mimic the interaction of stathmin with tubulin.  相似文献   

4.
Modifications of the D-piece carboxylic acid group of the hemiasterlin analog HTI-286 gave tubulin inhibitors which were potent cytotoxic agents in taxol resistant cell lines expressing P-glycoprotein. Amides derived from proline had potency comparable to HTI-286. Reduction of the carboxylic acid to ketones and alcohols or its conversion to acidic heterocycles also gave potent analogs. Synthetic modifications of the carboxylic acid could be carried out selectively using a wide range of synthetic reagents. Proline analog 3 was found to be effective in a human xenograft model in athymic mice.  相似文献   

5.
A series of tubulysin analogs in which one of the stereogenic centers of tubuphenylalanine was eliminated were synthesized. All compounds were tested for antiproliferative activity towards ovarian cancer cells and for inhibition of tubulin polymerization. The dimethyl analogs were generally more active than the desmethyl analogs, and four analogs have tubulin polymerization IC50 values similar to combretastatin A4 and the hemiasterlin analog HTI-286.  相似文献   

6.
Analogs of hemiasterlin (1) and HTI-286 (2), which contain various aromatic rings in the A segment, were synthesized as potential inhibitors of tubulin polymerization. The structure-activity relationships related to stereo- and regio-chemical effects of substituents on the aromatic ring in the A segment were studied. Analogs, which carry a meta-substituted phenyl ring in the A segment show comparable activity for inhibition of tubulin polymerization to 2, as well as in the cell proliferation assay using KB cells containing P-glycoprotein, compared to those of 1 and 2.  相似文献   

7.
Hemiasterlins are sponge-derived tripeptides that inhibit cell growth by depolymerizing existing microtubules and inhibiting microtubule assembly. Since hemiasterlins are poor substrates for P-glycoprotein, they are attractive candidates for cancer therapy and have been undergoing clinical trials. The basis of resistance to a synthetic analogue of hemiasterlin, HTI-286 (HTI), was examined in cell populations derived from ovarian carcinoma (A2780/1A9) cells selected in HTI-286. 1A9-HTI-resistant cells (1A9-HTI(R) series) were 57-89-fold resistant to HTI. Cross-resistance (3-186-fold) was observed to other tubulin depolymerizing drugs, with collateral sensitivity (2-14-fold) to tubulin polymerizing agents. Evaluation of the percentage of polymerized and soluble tubulin in 1A9 parental and 1A9-HTI(R) cells corroborated the HTI cytotoxicity data. At 22 degrees C or 37 degrees C, in the absence of any drug, the percentage of polymerized microtubules for each of the 1A9-HTI(R) populations was greater than that in the 1A9 parental cells, consistent with more stable microtubules. Furthermore, microtubules in the 1A9-HTI(R) populations were also more resistant to depolymerization at 4 degrees C and had more acetylated and detyrosinated (Glu-tubulin) alpha-tubulin, all characteristic of more stable microtubules. The 1A9-HTI(R) cell populations exhibited either a single nucleotide change in the M40 beta-tubulin isotype, S172A, or in two cell populations where no beta-tubulin mutation was detected, mutations in the Kalpha-1 alpha-tubulin isotype, S165P and R221H in one resistant cell population and I384V in another. Unlike reports of mutations resulting in reduced drug affinity, the experimental data and location of mutations are consistent with resistance to HTI-286 mediated by microtubule-stabilizing mutations in beta- or alpha-tubulin.  相似文献   

8.
The stereochemistry of the tubulin inhibitors taltobulin HTI-286 (2) and HTI-042 (3) was determined by utilizing the DPFGSE 1D NOE experiment. Single crystal X-ray diffraction analysis further confirmed the absolute configuration of these two compounds, which carry the (S,S,S)-configuration necessary for biological activity.  相似文献   

9.
Modifications of the B-segment of HTI-286 (2) produced a class of analogs incorporating heteroatom-substituents. The structure-activity relationship was studied. Analogs bearing methylsulfide and fluoride groups exhibited potency comparable to that of the parent compound HTI-286 and to paclitaxel in cytotoxicity assays against KB-3-1 cell lines. These analogs were more potent than paclitaxel against P-glycoprotein expressing KB-8-5 and KB-V1 cell lines. Several analogs showed strong inhibition of tubulin polymerization.  相似文献   

10.
The bacterial DNA ligase as a multiple domain protein is involved in DNA replication, repair and recombination. Its catalysis of ligation can be divided into three steps. To delineate the roles of amino acid residues in motif IV in ligation catalysis, site-directed mutants were constructed in a bacterial NAD^+-dependent DNA ligase from Thermus sp. TAK16D. It was shown that four conserved residues (D286, G287, V289 and K291) in motif IV had significant roles on the overall ligation. Under single turnover conditions, the observed apparent rates of D286E, G287A, V289I and K291R mutants were clearly reduced compared with that of WT ligase on both match and mismatch nicked substrates. The effects of D286E mutation on overall ligation may not only be ascribed to the third step. The G287A mutation has a major effect on the second step. The effects of V289I and K291R mutation on overall ligation are not on the third step, perhaps other aspects, such as conformation change of ligase protein in ligation catalysis, are involved. Moreover, the amino acid substitutions of above four residues were more sensitive on mismatch nicked substrate, indicating an enhanced ligation fidelity.  相似文献   

11.
Bai R  Durso NA  Sackett DL  Hamel E 《Biochemistry》1999,38(43):14302-14310
The sponge-derived antimitotic tripeptide hemiasterlin was previously shown to inhibit tubulin polymerization. We have now demonstrated that hemiasterlin resembles most other antimitotic peptides in noncompetitively inhibiting the binding of vinblastine to tubulin (apparent K(i) value, 7.0 microM), competitively inhibiting the binding of dolastatin 10 to tubulin (apparent K(i) value, 2.0 microM), stabilizing the colchicine binding activity of tubulin, inhibiting nucleotide exchange on beta-tubulin, and inducing the formation of tubulin oligomers that are stable to gel filtration in the absence of free drug, even at low drug concentrations. The tubulin oligomerization reaction induced by hemiasterlin was compared to the reactions induced by dolastatin 10 and cryptophycin 1. Like dolastatin 10, hemiasterlin induced formation of a tubulin aggregate that had the morphological appearance primarily of ring-like structures with a diameter of about 40 nm, while the morphology of the cryptophycin 1 aggregate consisted primarily of smaller rings (diameter about 30 nm). However, the hemiasterlin aggregate differed from the dolastatin 10 aggregate in that its formation was not associated with turbidity development, and the morphology of the hemiasterlin aggregate (as opposed to the dolastatin 10 aggregate) did not change greatly when microtubule-associated proteins were present (tight coils and pinwheels are observed with dolastatin 10 but not with hemiasterlin or cryptophycin 1). Opacification of tubulin-dolastatin 10 mixtures was inhibited by hemiasterlin at 22 degrees C and stimulated at 0 degrees C, while cryptophycin 1 was inhibitory at both reaction temperatures.  相似文献   

12.
Chemical and biological investigations of extracts from the sponge genus Auletta and two collections of Siphonochalina sp. have shown these organisms to be producers of the potent hemiasterlin class of antitumor agents. In addition to the previously known hemiasterlin (1) and hemiasterlin A (2), a new analogue, hemiasterlin C (3), was isolated and identified. The structures of 1 and 2 were assigned based on comparison to literature values, and 3 was identified on the basis of 1H NMR, 13C NMR, COSY, HSQC, and HMBC experiments. The cytotoxic and antitubulin activities of 1-3 were evaluated. In a comparative assay for inhibition of tubulin polymerization, the hemiasterlins were more potent than dolastatin 15 and equipotent with cryptophycin 1, but were somewhat less potent than dolastatin 10.  相似文献   

13.
Cytochrome c oxidase is the terminal electron acceptor in the respiratory chains of aerobic organisms and energetically couples the reduction of oxygen to water to proton pumping across the membrane. The mechanisms of proton uptake, gating, and pumping have yet to be completely elucidated at the molecular level for these enzymes. For Rhodobacter sphaeroides CytcO (cytochrome aa3), it appears as though the E286 side chain of subunit I is a branching point from which protons are shuttled either to the catalytic site for O2 reduction or to the acceptor site for pumped protons. Amide hydrogen-deuterium exchange mass spectrometry was used to investigate how mutation of this key branching residue to histidine (E286H) affects the structures and dynamics of four redox intermediate states. A functional characterization of this mutant reveals that E286H CytcO retains approximately 1% steady-state activity that is uncoupled from proton pumping and that proton transfer from H286 is significantly slowed. Backbone amide H-D exchange kinetics indicates that specific regions of CytcO, perturbed by the E286H mutation, are likely to be involved in proton gating and in the exit pathway for pumped protons. The results indicate that redox-dependent conformational changes around E286 are essential for internal proton transfer. E286H CytcO, however, is incapable of these specific conformational changes and therefore is insensitive to the redox state of the enzyme. These data support a model where the side chain conformation of E286 controls proton translocation in CytcO through its interactions with the proton gate, which directs the flow of protons either to the active site or to the exit pathway. In the E286H mutant, the proton gate does not function properly and the exit channel is unresponsive. These results provide new insight into the structure and mechanism of proton translocation by CytcO.  相似文献   

14.
The α1 Na/K-ATPase possesses both pumping and signaling functions. Using purified enzyme we found that the α1 Na/K-ATPase might interact with and regulate Src activity in a conformation-dependent manner. Here we further explored the importance of the conformational transition capability of α1 Na/K-ATPase in regulation of Src-related signal transduction in cell culture. We first rescued the α1-knockdown cells by wild-type rat α1 or α1 mutants (I279A and F286A) that are known to be defective in conformational transition. Stable cell lines with comparable expression of wild type α1, I279A, and F286A were characterized. As expected, the defects in conformation transition resulted in comparable degree of inhibition of pumping activity in the mutant-rescued cell lines. However, I279A was more effective in inhibiting basal Src activity than either the wild-type or the F286A. Although much higher ouabain concentration was required to stimulate Src in I279A-rescued cells, extracellular K+ was comparably effective in regulating Src in both control and I279A cells. In contrast, ouabain and extracellular K+ failed to produce detectable changes in Src activity in F286A-rescued cells. Furthermore, expression of either mutant inhibited integrin-induced activation of Src/FAK pathways and slowed cell spreading processes. Finally, the expression of these mutants inhibited cell growth, with I279A being more potent than that of F286A. Taken together, the new findings suggest that the α1 Na/K-ATPase may be a key player in dynamic regulation of cellular Src activity and that the capability of normal conformation transition is essential for both pumping and signaling functions of α1 Na/K-ATPase.  相似文献   

15.
The densin C-terminal domain can target Ca(2+)/calmodulin-dependent protein kinase IIα (CaMKIIα) in cells. Although the C-terminal domain selectively binds CaMKIIα in vitro, full-length densin associates with CaMKIIα or CaMKIIβ in brain extracts and in transfected HEK293 cells. This interaction requires a second central CaMKII binding site, the densin-IN domain, and an "open" activated CaMKII conformation caused by Ca(2+)/calmodulin binding, autophosphorylation at Thr-286/287, or mutation of Thr-286/287 to Asp. Mutations in the densin-IN domain (L815E) or in the CaMKIIα/β catalytic domain (I205/206K) disrupt the interaction. The amino acid sequence of the densin-IN domain is similar to the CaMKII inhibitor protein, CaMKIIN, and a CaMKIIN peptide competitively blocks CaMKII binding to densin. CaMKII is inhibited by both CaMKIIN and the densin-IN domain, but the inhibition by densin is substrate-selective. Phosphorylation of a model peptide substrate, syntide-2, or of Ser-831 in AMPA receptor GluA1 subunits is fully inhibited by densin. However, CaMKII phosphorylation of Ser-1303 in NMDA receptor GluN2B subunits is not effectively inhibited by densin in vitro or in intact cells. Thus, densin can target multiple CaMKII isoforms to differentially modulate phosphorylation of physiologically relevant downstream targets.  相似文献   

16.
Tzortzopoulos A  Török K 《Biochemistry》2004,43(21):6404-6414
The role of adenosine 5'-triphosphate (ATP) in the activation mechanism of alpha-Ca(2+)/calmodulin-dependent protein kinase II (alphaCaMKII) was investigated using the T286A non-autophosphorylatable mutant of alphaCaMKII. Characterization of the T286A-alphaCaMKII mutant revealed k(cat) = 0.06 +/- 0.02 s(-1) for the T286A mutant, a 6 (+/- 2)-fold lower value compared to wild-type alphaCaMKII with 100 microM smooth muscle myosin light chain (MLC) as substrate. MLC phosphorylation by the T286A mutant and wild-type alphaCaMKII was cooperative, with Hill coefficients 2.3 +/- 0.1 and 2.4 +/- 0.3, respectively. K(m) values for MLC were 96 +/- 28 microM with T286A-alphaCaMKII and 49 +/- 29 microM for wild-type alphaCaMKII. Thus, while the activity of alphaCaMKII was sensitive to mutation of the Thr(286) residue to Ala, the mechanisms of the wild-type and T286A mutant enzyme appeared similar. K(d) for Ca(2+)/calmodulin was 2-fold reduced to 40 nM compared to that of wild-type alphaCaMKII (75 nM). ATP induced a 9-fold stabilization of Ca(2+)/calmodulin binding to the T286A mutant enzyme. Fluorescence stopped-flow kinetic experiments revealed that two Ca(2+)/calmodulin-enzyme complexes were formed, the first, unaffected by ATP, with association and dissociation rate constants of 2 x 10(7) M(-1) s(-1) and 5 s(-1), respectively, containing calmodulin in extended conformation. The second complex, in which calmodulin adopted a compact conformation, was formed with association rate constant 3 x 10(6) M(-1) s(-1) and dissociation at 0.15 s(-1) in the absence and 0.015 s(-1) in the presence of ATP. These data show that ATP is involved in the activation mechanism by forming two classes of Ca(2+)/calmodulin.alphaCaMKII.ATP complex. It is likely that only one of the complexes is on the activation pathway.  相似文献   

17.
An activation switch in the ligand binding pocket of the C5a receptor   总被引:1,自引:0,他引:1  
Although agonists are thought to occupy binding pockets within the seven-helix core of serpentine receptors, the topography of these binding pockets and the conformational changes responsible for receptor activation are poorly understood. To identify the ligand binding pocket in the receptor for complement factor 5a (C5aR), we assessed binding affinities of hexapeptide ligands, each mutated at a single position, for seven mutant C5aRs, each mutated at a single position in the putative ligand binding site. In ChaW (an antagonist) and W5Cha (an agonist), the side chains at position 5 are tryptophan and cyclohexylalanine, respectively. Comparisons of binding affinities indicated that the hexapeptide residue at this position interacts with two C5aR residues, Ile-116 (helix III) and Val-286 (helix VII); in a C5aR model these two side chains point toward one another. Both the I116A and the V286A mutations markedly increased binding affinity of W5Cha but not that of ChaW. Moreover, ChaW, the antagonist hexapeptide, acted as a full agonist on the I116A mutant. These results argue that C5aR residues Ile-116 and Val-286 interact with the side chain at position 5 of the hexapeptide ligand to form an activation switch. Based on this and previous work, we present a docking model for the hexapeptide within the C5aR binding pocket. We propose that agonists induce a small change in the relative orientations of helices III and VII and that these helices work together to allow movement of helix VI away from the receptor core, thereby triggering G protein activation.  相似文献   

18.
19.
A molecular dynamics simulation of the Sm binding site from human U4 snRNA was undertaken to determine the conformational flexibility of this region and to identify RNA conformations that were important for binding of the Sm proteins. The RNA was fully-solvated (>9,000 water molecules) and charge neutralized by inclusion of potassium ions. A three nanosecond MD simulation was conducted using AMBER with long-range electrostatic forces considered using the particle mesh Ewald summation method. The initial model of the Sm binding site region had the central and 3' stem-loops that flanked the Sm site co-axial with one another, and with the single-stranded Sm binding site region ([I] conformation). During the course of the trajectory, the axes of the 3' stem-loop, and later the central stem-loop, became roughly orthogonal from their original anti-parallel orientation. As these conformational changes occurred, the snRNA adopted first an [L] conformation, and finally a [U] conformation. The [U] conformation was more stable than either the [I] or [L] conformations, and persisted for the final 1 ns of the trajectory. Analysis of the structure resulting from the MD simulations revealed the bulged nucleotide, U114, and the mismatched Ag91-G110 base pair provided distinctive structural features that may enhance Sm protein binding. Based on the results of the MD simulation and the available experimental data, we proposed a mechanism for the binding of the Sm protein sub-complexes to the snRNA. In this model, the D1/D2 and E/F/G Sm protein sub-complexes first bind the snRNA in the [U] conformation, followed by conformational re-arrangement to the [I] conformation and binding of the D3/B Sm protein sub-complex.  相似文献   

20.
Calcium/calmodulin-dependent protein kinase II (CaMKII) interprets information conveyed by the amplitude and frequency of calcium transients by a controlled transition from an autoinhibited basal intermediate to an autonomously active phosphorylated intermediate (De Koninck and Schulman, 1998). We used spin labelling and electron paramagnetic resonance spectroscopy to elucidate the structural and dynamic bases of autoinhibition and activation of the kinase domain of CaMKII. In contrast to existing models, we find that autoinhibition involves a conformeric equilibrium of the regulatory domain, modulating substrate and nucleotide access. Binding of calmodulin to the regulatory domain induces conformational changes that release the catalytic cleft, activating the kinase and exposing an otherwise inaccessible phosphorylation site, threonine 286. Autophosphorylation at Thr286 further disrupts the interactions between the catalytic and regulatory domains, enhancing the interaction with calmodulin, but maintains the regulatory domain in a dynamic unstructured conformation following dissociation of calmodulin, sustaining activation. These findings support a mechanistic model of the CaMKII holoenzyme grounded in a dynamic understanding of autoregulation that is consistent with a wealth of biochemical and functional data.  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