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 46 毫秒
1.
Cell surface calreticulin (CRT) binding to thrombospondin-1 (TSP1), regulates cell adhesion, migration, anoikis resistance, and collagen production. Due to the essential role of membrane microdomains in CRT-mediated focal adhesion disassembly, we previously studied the effect of raft-like bilayers on TSP1–CRT interactions with all-atom molecular dynamics (AAMD) simulations. However, the simulated systems of protein on the surface of the bilayer(s) in the explicit solvent are too large for long timescale AAMD simulations due to computational expense. In this study, we adopted a multiscale modeling approach of combining AAMD, coarse-grained molecule dynamics (CGMD), and reversed AAMD (REV AAMD) simulations to investigate the interactions of single CRT or of the TSP1–CRT complex with a membrane microdomain at microsecond timescale. Results showed that CRT conformational stabilization by binding of TSP1 in AAMD simulation was undetectable in CGMD simulation, but it was recovered in REV AAMD simulation. Similarly, interactions of the CRT N-domain and TSP1 with the membrane microdomain were lost in CGMD simulations but they were re-gained in the REV AAMD simulations. There was the higher coordination of the CRT P-domain in the TSP1–CRT complex with the lipid components of membrane microdomain compared to that of single CRT, which could directly affect the conformation of CRT and further mediate CRT recruitment of LDL receptor-related protein for signaling events. This study provides structural and molecular insights into TSP1–CRT interactions in a membrane microdomain environment and demonstrates the feasibility of using multiscale simulations to investigate the interactions between protein and membrane microdomains at a long timescale.  相似文献   

2.
The recognition of extracellular matrix components can be regulated by conformational changes that alter the activity of cell surface integrins. We now demonstrate that conformational regulation of the matrix glycoprotein thrombospondin-1 (TSP1) can also modulate its binding to an integrin receptor. F18 1G8 is a conformation-sensitive TSP1 antibody that binds weakly to soluble TSP1 in the presence of divalent cations. However, binding of the antibody to melanoma cells was strongly stimulated by adding exogenous TSP1 in the presence of calcium, suggesting that TSP1 undergoes a conformational change following its binding to the cell surface. This conformation was not induced by known cell surface TSP1 receptors, whereas binding of F18 was stimulated when TSP1 bound to fibronectin but not to heparin or fibrinogen. Conversely, binding of F18 to TSP1 enhanced TSP1 binding to fibronectin. Exogenous fibronectin also stimulated TSP1-dependent binding of F18 to melanoma cells. Binding of the fibronectin-TSP1 complex to melanoma cells was mediated by alpha4beta1 and alpha5beta1 integrins. Furthermore, binding to F18 or fibronectin strongly enhanced the adhesive activity of immobilized TSP1 for some cell types. This enhancement of adhesion was mediated by alpha3beta1 integrin and required that the alpha3beta1 integrin be in an active state. Fibronectin also enhanced TSP1 binding to purified alpha3beta1 integrin. Therefore, both fibronectin and the F18 antibody induce conformational changes in TSP1 that enhance the ability of TSP1 to be recognized by alpha3beta1 integrin. The conformational and functional regulation of TSP1 activity by fibronectin represents a novel mechanism for extracellular signal transduction.  相似文献   

3.
In this study, we examined the binding of soluble TSP1 (and ox-LDL) to CD36-transfected cells and the mechanisms by which immobilized TSP1 mediated attachment and haptotaxis (cell migration towards a substratum-bound ligand) of these transfected cells. CD36 cDNA transfection of NIH 3T3 cells clearly induced a dramatic increase in binding of both soluble [125I]-TSP1 and [125I]-ox-LDL to the surface of CD36-transfected cells, indicating that there was a gain of function with CD36 transfection in NIH 3T3 cells. Despite this gain of function, mock- and CD36-transfected NIH 3T3 cells attached and migrated to a similar extent on immobilized TSP1. An anti-TSP1 oligoclonal antibody inhibited CD36-transfected cell attachment to TSP1 while function blocking anti-CD36 antibodies, alone or in combination with heparin, did not. A series of fusion proteins encompassing cell-recognition domains of TSP1 was then used to delineate mechanisms by which NIH 3T3 cells adhere to TSP1. Although CD36 binds soluble TSP1 through a CSVTCG sequence located within type 1 repeats,18,19 CD36-transfected NIH 3T3 cells did not attach to immobilized type 1 repeats while they did adhere to the N-terminal, type 3 repeats (in an RGD-dependent manner) and the C-terminal domain of TSP1. Conversely, Bowes melanoma cells attached to type 1 repeats and the N- and C-terminal domains of TSP1. However, CD36 cDNA transfection of Bowes cells did not increase cell attachment to type 1 repeats compared to that observed with mock-transfected Bowes cells. Moreover, a function blocking anti-CSVTCG peptide antibody did not inhibit the attachment of mock- and CD36-transfected Bowes cells to type 1 repeats. It is suggested that CD36/TSP1 interaction does not occur upon cell–matrix adhesion and haptotaxis because TSP1 undergoes conformational changes that do not allow the exposure of the CD36 binding site. © 1998 John Wiley & Sons, Ltd.  相似文献   

4.
We describe here an alanine scanning mutational analysis of the Abeta(1-40) amyloid fibril monitored by fibril elongation thermodynamics derived from critical concentration values for fibril growth. Alanine replacement of most residues in the amyloid core region, residues 15-36, leads to destabilization of the elongation step, compared to wild-type, by about 1kcal/mol, consistent with a major role for hydrophobic packing in Abeta(1-40) fibril assembly. Where comparisons are possible, the destabilizing effects of Ala replacements are generally in very good agreement with the effects of Ala replacements of the same amino acid residues in an element of parallel beta-sheet in the small, globular protein Gbeta1. We utilize these Ala-WT DeltaDeltaG values to filter previously described Pro-WT DeltaDeltaG values, creating Pro-Ala DeltaDeltaG values that specifically assess the sensitivity of a sequence position, in the structural context of the Abeta fibril, to replacement by proline. The results provide a conservative view of the energetics of Abeta(1-40) fibril structure, indicating that positions 18-21, 25-26, and 32-33 within amyloid structure are particularly sensitive to the main-chain disrupting effects of Pro replacements. In contrast, residues 14-17, 22, 24, 27-31, and 34-39 are relatively insensitive to Pro replacements; most N-terminal residues were not tested. The results are discussed in terms of amyloid fibril structure and folding energetics, in particular focusing on how the data compare to those from other structural studies of Abeta(1-40) amyloid fibrils grown in phosphate-buffered saline at 37 degrees C under unstirred ("quiescent") conditions.  相似文献   

5.
A synthetic peptide containing amino acid residues 190-201 of thrombospondin-1 (TSP1) promoted adhesion of MDA-MB-435 breast carcinoma cells when immobilized and inhibited adhesion of the same cells to TSP1 when added in solution. Adhesion to this peptide was enhanced by a beta(1) integrin-activating antibody, Mn(2+), and insulin-like growth factor I and was inhibited by an alpha(3)beta(1) integrin function-blocking antibody. The soluble peptide inhibited adhesion of cells to the immobilized TSP1 peptide or spreading on intact TSP1 but at the same concentrations did not inhibit attachment or spreading on type IV collagen or fibronectin. Substitution of several residues in the TSP1 peptide with Ala residues abolished or diminished the inhibitory activity of the peptide in solution, but only substitution of Arg-198 completely inactivated the adhesive activity of the immobilized peptide. The essential residues for activity of the peptide as a soluble inhibitor are Asn-196, Val-197, and Arg-198, but flanking residues enhance the inhibitory activity of this core sequence, either by altering the conformation of the active sequence or by interacting with the integrin. This functional sequence is conserved in all known mammalian TSP1 sequences and in TSP1 from Xenopus laevis. The TSP1 peptide also inhibited adhesion of MDA-MB-435 cells to the laminin-1 peptide GD6, which contains a potential integrin-recognition sequence Asn-Leu-Arg and is derived from a similar position in a pentraxin module. Adhesion studies using recombinant TSP1 fragments also localized beta1 integrin-dependent adhesion to residues 175-242 of this region, which contain the active sequence.  相似文献   

6.
Interactions of heparin with intact human thrombospondin-1 (TSP1) and with two heparin-binding fragments of TSP1 were characterized using chemically modified heparins, a vascular heparan sulfate proteoglycan, and a series of heparin oligosaccharides prepared by partial deaminative cleavage. The avidity of TSP1 binding increased with oligosaccharide size, with plateaus at 4 to 6 and at 8 to 10 monosaccharide units. The dependence on oligosaccharide size for binding to the recombinant amino-terminal heparin-binding domain of TSP1 was the same as that of the intact TSP1 molecule but differed from that of a synthetic heparin-binding peptide from the type 1 repeats, suggesting that the interaction between intact TSP1 and heparin is primarily mediated by the amino-terminal domain. Based on activities of chemically modified heparins, binding to TSP1 depended primarily on 2-N- and 6-O-sulfation of glucosamine and to a lesser degree on 2,3-O-sulfation and the carboxyl residues of the uronic acids. In contrast, all of these modifications were required for binding of heparin to the type 1 repeat peptides. Affinity purification of heparin octasaccharides on immobilized TSP1 type 1 repeat peptides revealed a preference for oligosaccharides containing the disaccharide sequence IdoA(2-OSO(3))alpha1-4-GlcNS(6-OSO(3)). Binding of these oligosaccharides to the peptide required the Trp residues. These data demonstrate that the heparin-binding specificities of intact TSP1 and peptides from the type 1 repeats overlap with that of basic fibroblast growth factor (FGF2) and are consistent with the ability of these TSP1-derived molecules to inhibit FGF2-stimulated angiogenesis.  相似文献   

7.
Aspartate decarboxylase, which is translated as a pro-protein, undergoes intramolecular self-cleavage at Gly24-Ser25. We have determined the crystal structures of an unprocessed native precursor, in addition to Ala24 insertion, Ala26 insertion and Gly24-->Ser, His11-->Ala, Ser25-->Ala, Ser25-->Cys and Ser25-->Thr mutants. Comparative analyses of the cleavage site reveal specific conformational constraints that govern self-processing and demonstrate that considerable rearrangement must occur. We suggest that Thr57 Ogamma and a water molecule form an 'oxyanion hole' that likely stabilizes the proposed oxyoxazolidine intermediate. Thr57 and this water molecule are probable catalytic residues able to support acid-base catalysis. The conformational freedom in the loop preceding the cleavage site appears to play a determining role in the reaction. The molecular mechanism of self-processing, presented here, emphasizes the importance of stabilization of the oxyoxazolidine intermediate. Comparison of the structural features shows significant similarity to those in other self-processing systems, and suggests that models of the cleavage site of such enzymes based on Ser-->Ala or Ser-->Thr mutants alone may lead to erroneous interpretations of the mechanism.  相似文献   

8.
Ash MR  Chong LX  Maher MJ  Hinds MG  Xiao Z  Wedd AG 《Biochemistry》2011,50(43):9237-9247
The bacterium Cupriavidus metallidurans CH34 is resistant to high environmental concentrations of many metal ions. Upon copper challenge, it upregulates the periplasmic protein CopK (8.3 kDa). The function of CopK in the copper resistance response is ill-defined, but CopK demonstrates an intriguing cooperativity: occupation of a high-affinity Cu(I) binding site generates a high-affinity Cu(II) binding site, and the high-affinity Cu(II) binding enhances Cu(I) binding. Native CopK and targeted variants were examined by chromatographic, spectroscopic, and X-ray crystallographic probes. Structures of two distinct forms of Cu(I)Cu(II)-CopK were defined, and structural changes associated with occupation of the Cu(II) site were demonstrated. In solution, monomeric Cu(I)Cu(II)-CopK features the previously elucidated Cu(I) site in Cu(I)-CopK, formed from four S(δ) atoms of Met28, -38, -44, and -54 (site 4S). Binding of Cu(I) to apo-CopK induces a conformational change that releases the C-terminal β-strand from the β-sandwich structure. In turn, this allows His70 and N-terminal residues to form a large loop that includes the Cu(II) binding site. In crystals, a polymeric form of Cu(I)Cu(II)-CopK displays a Cu(I) site defined by the S(δ) atoms of Met26, -38, and -54 (site 3S) and an exogenous ligand (modeled as H(2)O) and a Cu(II) site that bridges dimeric CopK molecules. The 3S Cu(I) binding mode observed in crystals was demonstrated in solution in protein variant M44L where site 4S is disabled. The intriguing copper binding chemistry of CopK provides molecular insight into Cu(I) transfer processes. The adaptable nature of the Cu(I) coordination sphere in methionine-rich clusters allows copper to be relayed between clusters during transport across membranes in molecular pumps such as CusA and Ctr1.  相似文献   

9.
The creatine transporter (CRT) is a member of a large family of sodium-dependent neurotransmitter and amino acid transporters. The CRT is closely related to the gamma-aminobutyric acid (GABA) transporter, GAT-1, yet GABA is not an effective substrate for the CRT. The high resolution structure of a prokaryotic homologue, LeuT has revealed precise details of the substrate binding site for leucine (Yamashita, A., Singh, S. K., Kawate, T., Jin, Y., and Gouaux, E. (2005) Nature 437, 215-223). We have now designed mutations based on sequence comparisons of the CRT with GABA transporters and the LeuT structural template in an attempt to alter the substrate specificity of the CRT. Combinations of two or three amino acid substitutions at four selected positions resulted in the loss of creatine transport activity and gain of a specific GABA transport function. GABA transport by the "gain of function" mutants was sensitive to nipecotic acid, a competitive inhibitor of GABA transporters. Our results show LeuT to be a good structural model to identify amino acid residues involved in the substrate and inhibitor selectivity of eukaryotic sodium-dependent neurotransmitter and amino acid transporters. However, modification of the binding site alone appears to be insufficient for efficient substrate translocation. Additional residues must mediate the conformational changes required for the diffusion of substrate from the binding site to the cytoplasm.  相似文献   

10.
Ras GTPases function as binary switches in signaling pathways controlling cell growth and differentiation. The guanine nucleotide exchange factor Sos mediates the activation of Ras in response to extracellular signals. We have previously solved the crystal structure of nucleotide-free Ras in complex with the catalytic domain of Sos (Boriack-Sjodin, P. A., Margarit, S. M., Bar-Sagi, D., and Kuriyan, J. (1998) Nature 394, 337-343). The structure demonstrates that Sos induces conformational changes in two loop regions of Ras known as switch 1 and switch 2. In this study, we have employed site-directed mutagenesis to investigate the functional significance of the conformational changes for the catalytic function of Sos. Switch 2 of Ras is held in a very tight embrace by Sos, with almost every external side chain coordinated by Sos. Mutagenesis of contact residues at the switch 2-Sos interface shows that only a small set of side chains affect binding, with the most important contact being mediated by tyrosine 64, which is buried in a hydrophobic pocket of Sos in the Ras.Sos complex. Substitutions of Ras and Sos side chains that are inserted into the Mg(2+)- and nucleotide phosphate-binding site of switch 2 (Ras Ala(59) and Sos Leu(938) and Glu(942)) have no effect on the catalytic function of Sos. These results indicate that the interaction of Sos with switch 2 is necessary for tight binding, but is not the critical driving force for GDP displacement. The structural distortion of switch 1 induced by Sos is mediated by a small number of specific contacts between highly conserved residues on both Ras and Sos. Mutations of a subset of these residues (Ras Tyr(32) and Tyr(40)) result in an increase in the intrinsic rate of nucleotide dissociation from Ras and impair the binding of Ras to Sos. Based on this analysis, we propose that the interactions of Sos with the switch 1 and switch 2 regions of Ras have distinct functional consequences: the interaction with switch 2 mediates the anchoring of Ras to Sos, whereas the interaction with switch 1 leads to disruption of the nucleotide-binding site and GDP dissociation.  相似文献   

11.
Crotonase from Clostridium acetobutylicum (CaCRT) is an enzyme that catalyzes the dehydration of 3-hydroxybutyryl-CoA to crotonyl-CoA in the n-butanol biosynthetic pathway. To investigate the molecular mechanism underlying n-butanol biosynthesis, we determined the crystal structures of the CaCRT protein in apo- and acetoacetyl-CoA bound forms. Similar to other canonical crotonase enzymes, CaCRT forms a hexamer by the dimerization of two trimers. A crystal structure of CaCRT in complex with acetoacetyl-CoA revealed that Ser69 and Ala24 to be signature residues of CaCRT, which results in a distinct ADP binding mode wherein the ADP moiety is bound at a different position compared with other crotonases. We also revealed that the substrate specificity of crotonase enzymes is determined by both the structural feature of the α3 helix region and the residues contributing the enoyl-CoA binding pocket. A tight formed α3 helix and two phenylalanine residues, Phe143 and Phe233, aid CaCRT to accommodate crotonyl-CoA as the substrate. The key residues involved in substrate binding, enzyme catalysis and substrate specificity were confirmed by site-directed mutagenesis.  相似文献   

12.
Lim KH  Kim YK  Chang YT 《Biochemistry》2007,46(47):13523-13532
Transition-metal ions (Cu2+ and Zn2+) play critical roles in the Abeta plaque formation. However, precise roles of the metal ions in the Abeta amyloidogenesis have been controversial. In this study, the molecular mechanism of the metal-induced Abeta oligomerization was investigated with extensive metal ion titration NMR experiments. Upon additions of the metal ions, the N-terminal region (1-16) of the Abeta (1-40) peptide was selectively perturbed. In particular, polar residues 4-8 and 13-15 were more strongly affected by the metal ions, suggesting that those regions may be the major binding sites of the metal ions. The NMR signal changes of the N-terminal region were dependent on the peptide concentrations (higher peptide concentrations resulted in stronger signal changes), suggesting that the metal ions facilitate the intermolecular contact between the Abeta peptides. The Abeta (1-40) peptides (>30 microM) were eventually oligomerized even at low temperature (3 degrees C), where the Abeta peptides are stable as monomeric forms without the metal ions. The real-time oligomerization process was monitored by 1H/15N HSQC NMR experiments, which provided the first residue-specific structural transition information. Hydrophobic residues 12-21 initially underwent conformational changes due to the intermolecular interactions. After the initial structural rearrangements, the C-terminal residues (32-40) readjusted their conformations presumably for effective oligomerization. Similar structural changes of the metal-free Abeta (1-40) peptides were also observed in the presence of the preformed oligomers, suggesting that the conformational transitions may be the general molecular mechanism of the Abeta (1-40) amyloidogenesis.  相似文献   

13.
The extracellular carbohydrate-binding domain of the Type I transmembrane receptor CD44 is known to undergo affinity switching, where change in conformation leads to enhanced binding of its carbohydrate ligand hyaluronan. Separate x-ray crystallographic and NMR experiments have led to competing explanations, with the former supporting minor conformational changes at the binding site and the latter a major order-to-disorder unfolding transition distant from the binding site. Here, all-atom explicit-solvent molecular dynamics studies employing adaptive biasing force sampling revealed a substantial favorable free-energy change associated with contact formation between the Arg41 side chain and hyaluronan at the binding site, independent of whether the distant site was ordered or disordered. Analogous computational experiments on Arg41Ala mutants showed loss of this favorable free-energy change, consistent with existing experimental data. More provocatively, the simulation data revealed the molecular mechanism by which the order-to-disorder transition enhances hyaluronan binding: in the disordered state, a number of basic residues gain sufficient conformational freedom—lacking in the ordered state—to spontaneously form side-chain contacts with hyaluronan. Mutation of these residues to Ala had been known to decrease binding affinity, but there had previously been no structural explanation, given their lack of proximity to the carbohydrate-binding site in existing structures of the complex.  相似文献   

14.
The extracellular carbohydrate-binding domain of the Type I transmembrane receptor CD44 is known to undergo affinity switching, where change in conformation leads to enhanced binding of its carbohydrate ligand hyaluronan. Separate x-ray crystallographic and NMR experiments have led to competing explanations, with the former supporting minor conformational changes at the binding site and the latter a major order-to-disorder unfolding transition distant from the binding site. Here, all-atom explicit-solvent molecular dynamics studies employing adaptive biasing force sampling revealed a substantial favorable free-energy change associated with contact formation between the Arg41 side chain and hyaluronan at the binding site, independent of whether the distant site was ordered or disordered. Analogous computational experiments on Arg41Ala mutants showed loss of this favorable free-energy change, consistent with existing experimental data. More provocatively, the simulation data revealed the molecular mechanism by which the order-to-disorder transition enhances hyaluronan binding: in the disordered state, a number of basic residues gain sufficient conformational freedom—lacking in the ordered state—to spontaneously form side-chain contacts with hyaluronan. Mutation of these residues to Ala had been known to decrease binding affinity, but there had previously been no structural explanation, given their lack of proximity to the carbohydrate-binding site in existing structures of the complex.  相似文献   

15.
Sorting of the prohormone POMC to the regulated secretory pathway necessitates the binding of a sorting signal to a sorting receptor, identified as membrane carboxypeptidase E (CPE). The sorting signal, located at the N terminus of POMC consists of two acidic (Asp10, Glu14) and two hydrophobic (Leu11, Leu18) residues exposed on the surface of an amphipathic loop. In this study, molecular modeling of CPE predicted that the acidic residues in the POMC-sorting signal bind specifically to two basic residues, Arg255 and Lys260, present in a loop unique to CPE, compared with other carboxypeptidases. To test the model, these two residues on CPE were mutated to Ser or Ala, followed by baculovirus expression of the mutant CPEs in Sf9 cells. Sf9 cell membranes containing CPE mutants with either Arg255 or Lys260, or both residues substituted, showed no binding of [125I]N-POMC1-26 (which contains the POMC-sorting signal motif), proinsulin, or proenkephalin. In contrast, substitution of an Arg147 to Ala147 at a substrate-binding site, Arg259 to Ala259 and Ser202 to Pro202, in CPE did not affect the level of [125I]N-POMC1-26 binding when compared with-wild type CPE. Furthermore, mutation of the POMC-sorting signal motif (Asp10, Leu11, Glu14, Leu18) eliminated binding to wild-type CPE. These results indicate that the sorting signal of POMC, proinsulin, and proenkephalin specifically interacts with Arg255 and Lys260 at a novel binding site, independent of the active site on CPE.  相似文献   

16.
Phenylalanine hydroxylase (PAH) is generally considered to undergo a large and reversible conformational transition upon l-Phe binding, which is closely linked to the substrate-induced catalytic activation of this hysteretic enzyme. Recently, several crystallographically solvent-exposed hinge-bending regions including residues 31-34, 111-117, 218-226, and 425-429 have been defined/predicted to be involved in the intra-protomer propagation of the substrate-triggered molecular motions generated at the active site. On this basis, single-site mutagenesis of key residues in these regions of the human PAH tetramer was performed in the present study, and their functional impact was measured by steady-state kinetics and the global conformational transition as assessed by surface plasmon resonance and intrinsic tryptophan fluorescence spectroscopy. A strong correlation (r(2) = 0.93-0.96) was observed between the l-Phe-induced global conformational transition and V(max) values for wild-type human PAH and the mutant forms K113P, N223D, N426D, and N32D, in contrast to the substitution T427P, which resulted in a tetrameric form with no kinetic cooperativity. Furthermore, the flexible intra-domain linker region (residues 31-34) seems to be involved in a more local conformational change, and the biochemical/biophysical properties of the G33A/G33V mutant forms support a key function of this residue in the positioning of the autoregulatory sequence (residues 1-30) and thus in the regulation of the solvent and substrate access to the active site. The mutant forms revealed a variably reduced global conformational stability compared with wild-type human PAH, as measured by thermal denaturation and limited proteolysis.  相似文献   

17.
An NMR study has been made of porcine pancreatic phospholipase A2 (PLA) in three environments: free in solution, in a binary complex with dodecylphosphocholine micelles, and in a ternary complex with a micelle and the substrate-like inhibitor (R)-1-octyl-2-(N-dodecanoylamino)-2-deoxyglycero-3-phosph oglycol. 1H and 15N chemical shifts, amide exchange rates, and NOE intensities are compared for the enzyme in different environments. From these data, structural differences are found for the N-terminal part, the end of the surface loop at residues Tyr69 and Thr70, and the active site residue His48, and also for the Ca-binding loop (residues 28-32). Specifically, when binding to a micelle, the side chains of residues Ala1, Trp3, and Tyr69, as well as all protons of Thr70, are found to be closer together. After subsequent introduction of the competitive inhibitor, further changes are found for these residues. The N-terminus is flexible in PLA free in solution, in contrast with the crystal structures where it adopts an alpha-helical conformation. According to the NMR data, this helix is rigidly formed only in the ternary complex. Furthermore, in the ternary complex, the N-terminal amino group and the exchangeable hydrogen at N3 of the ring of His48 are observed. We propose that PLA is activated in two steps. An initial conformational change occurs upon binding to a micellar interface. The catalytically active conformation of the enzyme, which has an extensive network of hydrogen bonds, is formed only when binding a substrate or competitive inhibitor at a lipid-water interface.  相似文献   

18.
Thrombospondin-1 (TSP1) is a large modular matrix protein containing three identical disulfide-linked 180-kD chains that inhibits neovascularization in vivo (Good et al., 1990). To determine which of the structural motifs present in the 180-kD TSP1 polypeptide mediate the anti-angiogenic activity, a series of protease-generated fragments were tested using several in vitro and in vivo assays that reflect angiogenic activity. The majority of the anti-angiogenic activity of TSP1 resides in the central 70-kD stalk region which alone could block neovascularization induced by bFGF in the rat cornea in vivo and inhibit both migration in a modified Boyden chamber and [3H]thymidine incorporation stimulated by bFGF in cultured capillary endothelial cells. Although TSP1 has been shown to bind active TGF beta 1, this cytokine could not account for the inhibitory effects of the stalk region of TSP1 on cultured endothelial cells. Peptides and truncated molecules were used to further localize inhibitory activity to two domains of the central stalk, the procollagen homology region and the properdin-like type 1 repeats. Trimeric recombinant TSP1 containing NH2- terminal sequences truncated after the procollagen-like module inhibited endothelial cell migration in vitro and corneal neovascularization in vivo whereas trimeric molecules truncated before this domain were inactive as was the NH2-terminal heparin-binding domain that is present in both recombinant molecules. A series of peptides from the procollagen-like region, the smallest of which consisted of residues 303-309 of TSP1, inhibited angiogenesis in vivo in the rat cornea and the migration of endothelial cells in vitro. A 19- residue peptide containing these sequences blocked vessel formation in the granulation tissue invading a polyvinyl sponge implanted into the mouse. Nineteen residue peptides derived from two of the three type 1 repeats present in the intact TSP1 molecule blocked neovascularization in vivo in the rat cornea and inhibited the migration of cultured endothelial cells with ED50's of 0.6-7 microM. One of these peptides, containing residues 481-499 of TSP1, also inhibited vessel formation in granulation tissue invading sponges in vivo. These results suggest that the large TSP1 molecule employs at least two different structural domains and perhaps two different mechanisms to accomplish a single physiological function, the inhibition of neovascularization. The definition of short peptides from each of these domains that are able to block the angiogenic process may be of use in designing targeted inhibitors of the pathological neovascularization that underlies many diseases.  相似文献   

19.
Li D  Xiao Y  Xu X  Xiong X  Lu S  Liu Z  Zhu Q  Wang M  Gu X  Liang S 《The Journal of biological chemistry》2004,279(36):37734-37740
Hainantoxin-IV (HNTX-IV) can specifically inhibit the neuronal tetrodotoxin-sensitive sodium channels and defines a new class of depressant spider toxin. The sequence of native HNTX-IV is ECLGFGKGCNPSNDQCCKSSNLVCSRKHRWCKYEI-NH(2). In the present study, to obtain further insight into the primary and tertiary structural requirements of neuronal sodium channel blockers, we determined the solution structure of HNTX-IV as a typical inhibitor cystine knot motif and synthesized four mutants designed based on the predicted sites followed by structural elucidation of two inactive mutants. Pharmacological studies indicated that the S12A and R26A mutants had activities near that of native HNTX-IV, while K27A and R29A demonstrated activities reduced by 2 orders of magnitude. (1)H MR analysis showed the similar molecular conformations for native HNTX-IV and four synthetic mutants. Furthermore, in the determined structures of K27A and R29A, the side chains of residues 27 and 29 were located in the identical spatial position to those of native HNTX-IV. These results suggested that residues Ser(12), Arg(26), Lys(27), and Arg(29) were not responsible for stabilizing the distinct conformation of HNTX-IV, but Lys(27) and Arg(29) were critical for the bioactivities. The potency reductions produced by Ala substitutions were primarily due to the direct interaction of the essential residues Lys(27) and Arg(29) with sodium channels rather than to a conformational change. After comparison of these structures and activities with correlated toxins, we hypothesized that residues Lys(27), Arg(29), His(28), Lys(32), Phe(5), and Trp(30) clustered on one face of HNTX-IV were responsible for ligand binding.  相似文献   

20.
Synthetic nonbasic peptides based on the type I repeats of thrombospondin (TSP) and four peptides corresponding to the predicted basic clusters in lipoprotein lipase (LPL) have been analyzed for heparin binding. In the present report we examine the structural requirement for the binding of these peptides to heparin-Sepharose column. The peptide containing the sequence Phe-Ser-Trp-Ser-Asp-Trp-Trp-Ser (residues 388–395 in lipoprotein lipase, which include the consensus TSP type I sequence) showed strong binding to heparin. Both the first and second Trp residues in this sequence were essential for tight heparin binding. Substitution of either of the Trp residues by an Ala resulted in the complete loss of heparin binding. The peptides representing the four basic cluster regions of lipoprotein lipase showed variable heparin binding. Strong retention was observed for peptides representing cluster 1 (residues 261–287) and cluster 3 (residues 147–151) peptides followed by cluster 2 (residues 290–302) peptide. A peptide corresponding to LPL cluster 4 (residues 405–414) did not show binding to heparin column. The present study confirms the presence of specific heparin-binding sites in LPL. Furthermore, this study also demonstrates the potential use of synthetic peptides to investigate the interaction between peptides and heparin as an alternative approach to site-directed mutagenesis in selected regions of large protein molecules. The affinity of these peptides toward heparin can be explored to block molecular interactions at these specific sites or to carry and deliver other coupled molecules at the site(s) of attachment of these peptides for therapeutic applications.  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