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 31 毫秒
1.
A chitooligosaccharide specific lectin (Luffa acutangula agglutinin) has been purified from the exudate of ridge gourd fruits by affinity chromatography on soybean agglutinin-glycopeptides coupled to Sepharose-6B. The affinity purified lectin was found homogeneous by polyacrylamide gel electrophoresis, in sodium dodecyl sulphate-polyacrylamide gels, by gel filtration on Sephadex G-100 and by sedimentation velocity experiments. The relative molecular weight of this lectin is determined to be 48,000 ±1,000 by gel chromatography and sedimentation equilibrium experiments. The sedimentation coefficient (S20, w) was obtained to be 4.06 S. The Stokes’ radius of the protein was found to be 2.9 nm by gel filtration. In sodium dodecyl sulphate-polyacrylamide gel electrophoresis the lectin gave a molecular weight of 24,000 in the presence as well as absence of 2-mercaptoethanol. The subunits in this dimeric lectin are therefore held by non-covalent interactions alone. The lectin is not a glycoprotein and circular dichroism spectral studies indicate that this lectin has 31% α-helix and no β-sheet. The lectin is found to bind specifically to chitooligosaccharides and the affinity of the lectin increases with increasing oligosaccharide chain length as monitored by near ultra-violet-circular dichroism and intrinsic fluorescence titration. The values of ΔG, ΔH and ΔS for the binding process showed a pronounced dependence on the size of the oligosaccharide. The values for both ΔH and ΔS show a significant increase with increase in the oligosaccharide chain length showing that the binding of higher oligomers is progressively more favoured thermodynamically than chitobiose itself. The thermodynamic data is consistent with an extended binding site in the lectin which accommodates a tetrasaccharide. Based on the thermodynamic data, blue shifts and fluorescence enhancement, spatial orientation of chitooligosaccharides in the combining site of the lectin is assigned.  相似文献   

2.
A key step in the rational design of new RNA binding small molecules necessitates a complete elucidation of the molecular aspects of the binding of existing molecules to RNA structures. This work focuses towards the understanding of the interaction of a DNA intercalator, quinacrine and a minor groove binder 4′,6-diamidino-2-phenylindole (DAPI) with the right handed Watson–Crick base paired A-form and the left-handed Hoogsteen base paired HL-form of poly(rC)·poly(rG) evaluated by multifaceted spectroscopic and viscometric techniques. The energetics of their interaction has also been elucidated by isothermal titration calorimetry. Results of this study converge to suggest that (i) quinacrine intercalates to both A-form and HL-form of poly(rC)·poly(rG); (ii) DAPI shows both intercalative and groove-binding modes to the A-form of the RNA but binds by intercalative mode to the HL-form. Isothermal calorimetric patterns of quinacrine binding to both the forms of RNA and of DAPI binding to the HL-form are indicative of single binding while the binding of DAPI to the A-form reveals two kinds of binding. The binding of both the drugs to both conformations of RNA is exothermic; while the binding of quinacrine to both conformations and DAPI to the A-form (first site) is entropy driven, the binding of DAPI to the second site of A-form and HL-conformation is enthalpy driven. Temperature dependence of the binding enthalpy revealed that the RNA–ligand interaction reactions are accompanied by small heat capacity changes that are nonetheless significant. We conclude that the binding affinity characteristics and energetics of interaction of these DNA binding molecules to the RNA conformations are significantly different and may serve as data for the development of effective structure selective RNA-based antiviral drugs.  相似文献   

3.
A galactose-binding lectin was isolated in electrophoretically pure form from the seeds of the snake gourd,Trichosanthes anguina, by affinity chromatography on an immobilised lactose column, as well as on a cross-linkedGuar Gum column. The lectin agglutinates native erythrocytes of human A, B and 0 phenotypes and of rabbit, rat and mouse. The molecular mass of the lectin, as estimated bySephadex G-200 gel chromatography, is 49 kDa. Sodium dodecyl sulphate-polyacrylamide gel electrophoresis, after reduction with β-mercaptoethanol, revealed two polypeptide chains linked by disulphide bonds in the lectin molecule. It contains no covalently linked sugars. Amino acid analysis of the lectin revealed a high content of acidic amino acids, relatively lower proportion of basic amino acids and traces of cysteine and methionine. The lectin has good thermal stability, and is inactivated when oxidised by metaperiodate.  相似文献   

4.
Small molecule inhibitors against chitinases have potential applications as pesticides, fungicides, and antiasthmatics. Here, we report that a series of fully deacetylated chitooligosaccharides (GlcN)2–7 can act as inhibitors against the insect chitinase OfChtI, the human chitinase HsCht, and the bacterial chitinases SmChiA and SmChiB with IC50 values at micromolar to millimolar levels. The injection of mixed (GlcN)2–7 into the fifth instar larvae of the insect Ostrinia furnacalis resulted in 85% of the larvae being arrested at the larval stage and death after 10 days, also suggesting that (GlcN)2–7 might inhibit OfChtI in vivo. Crystal structures of the catalytic domain of OfChtI (OfChtI-CAD) complexed with (GlcN)5,6 were obtained at resolutions of 2.0 Å. These structures, together with mutagenesis and thermodynamic analysis, suggested that the inhibition was strongly related to the interaction between the −1 GlcN residue of the inhibitor and the catalytic Glu148 of the enzyme. Structure-based comparison showed that the fully deacetylated chitooligosaccharides mimic the substrate chitooligosaccharides by binding to the active cleft. This work first reports the inhibitory activity and proposed inhibitory mechanism of fully deacetylated chitooligosaccharides. Because the fully deacetylated chitooligosaccharides can be easily derived from chitin, one of the most abundant materials in nature, this work also provides a platform for developing eco-friendly inhibitors against chitinases.  相似文献   

5.
Sclerotium rolfsii lectin (SRL), a secretory protein from the soil borne phytopathogenic fungus Sclerotium rolfsii, has shown in our previous studies to bind strongly to the oncofetal Thomson-Friedenreich carbohydrate (Galβ1-3GalNAc-ser/thr, T or TF) antigen. TF antigen is widely expressed in many types of human cancers and the strong binding of SRL toward such a cancer-associated carbohydrate structure led us to characterize the carbohydrate binding specificity of SRL. Glycan array analysis, which included 285 glycans, shows exclusive binding of SRL to the O-linked mucin type but not N-linked glycans and amongst the mucin type O-glycans, lectin recognizes only mucin core 1, core 2 and weakly core 8 but not to other mucin core structures. It binds with high specificity to “α-anomers” but not the “β-anomers” of the TF structure. The axial C4-OH group of GalNAc and C2-OH group of Gal is both essential for SRL interaction with TF disaccharide, and substitution on C3 of galactose by sulfate or sialic acid or N-acetylglucosamine, significantly enhances the avidity of the lectin. SRL differs in its binding to TF structures compared to other known TF-binding lectins such as the Arachis hypogea (peanut) agglutinin, Agaricus bisporus (mushroom) lectin, Jackfruit, Artocarpus integrifolia (jacalin) and Amaranthus caudatus (Amaranthin) lectin. Thus, SRL has unique carbohydrate-binding specificity toward TF-related O-linked carbohydrate structures. Such a binding specificity will make this lectin a very useful tool in future structural as well as functional analysis of the cellular glycans in cancer studies.  相似文献   

6.
Owing to the use of porphyrins in photodynamic therapy for the treatment of malignant tumors, and the preferential interaction of lectins with tumor cells, studies on lectin-porphyrin interaction are of significant interest. In this study, the interaction of several free-base and metalloporphyrins with Momordica charantia (bitter gourd) lectin (MCL) was investigated by absorption spectroscopy. Difference absorption spectra revealed that significant changes occur in the Soret band region of the porphyrins on binding to MCL. These changes were monitored to obtain association constants (Ka) and stoichiometry of binding. The tetrameric MCL binds four porphyrin molecules, and the stoichiometry was unaffected by the presence of the specific sugar, lactose. In addition, the agglutination activity of MCL was unaffected by the presence of the porphyrins used in this study, clearly indicating that porphyrin and carbohydrate ligands bind at different sites. Both cationic and anionic porphyrins bind to the lectin with comparable affinity (Ka =10(3)-10(5) m(-1)). The thermodynamic parameters associated with the interaction of several porphyrins, obtained from the temperature dependence of the Ka values, were found to be in the range: DeltaH degrees = -98.1 to -54.4 kJ.mol(-1) and DeltaS degrees =-243.9 to -90.8 J.mol(-1).K(-1). These results indicate that porphyrin binding to MCL is governed by enthalpic forces and that the contribution from binding entropy is negative. Enthalpy-entropy compensation was observed in the interaction of different porphyrins with MCL, underscoring the role of water structure in the overall binding process. Analysis of CD spectra of MCL indicates that this protein contains about 13%alpha-helix, 36%beta-sheet, 21%beta-turn, and the rest unordered structures. Binding of porphyrins does not significantly alter the secondary and tertiary structures of MCL.  相似文献   

7.
Four crystal structures of human YKL-39 were solved in the absence and presence of chitooligosaccharides. The structure of YKL-39 comprises a major (β/α)8 triose-phosphate isomerase barrel domain and a small α + β insertion domain. Structural analysis demonstrates that YKL-39 interacts with chitooligosaccharides through hydrogen bonds and hydrophobic interactions. The binding of chitin fragments induces local conformational changes that facilitate tight binding. Compared with other GH-18 members, YKL-39 has the least extended chitin-binding cleft, containing five subsites for sugars, namely (−3)(−2)(−1)(+1)(+2), with Trp-360 playing a prominent role in the sugar-protein interactions at the center of the chitin-binding cleft. Evaluation of binding affinities obtained from isothermal titration calorimetry and intrinsic fluorescence spectroscopy suggests that YKL-39 binds to chitooligosaccharides with Kd values in the micromolar concentration range and that the binding energies increase with the chain length. There were no significant differences between the Kd values of chitopentaose and chitohexaose, supporting the structural evidence for the five binding subsite topology. Thermodynamic analysis indicates that binding of chitooligosaccharide to YKL-39 is mainly driven by enthalpy.  相似文献   

8.
The PAK2/βPIX/GIT1 (p21-activated kinase 2/PAK-interacting exchange factor-β/G protein-coupled receptor kinase-interactor 1) complex has been shown to distribute to both membrane ruffles and focal adhesions of cells, where it plays an important role in regulating focal adhesion turnover. However, the detailed mechanism underlying this regulation is largely unknown. We previously reported that MYO18Aα interacts via its carboxyl terminus with the PAK2/βPIX/GIT1 complex through direct binding to βPIX, and that knockdown of MYO18Aα in epithelial cells causes accumulation of the complex in focal adhesions and decreased cell migration ability (Hsu et al., 2010). The current study characterized the detailed MYO18Aα–βPIX interaction mechanism and the biological significance of this interaction. We found that deletion of the carboxyl-terminal globular domain of MYO18Aα profoundly altered the cellular localization of βPIX and inhibited cell migration. βPIX interacts through its most carboxyl-terminus, PAWDETNL (639–646), with MYO18Aα and partially colocalized with MYO18Aα in membrane ruffles of cells, whereas βPIX1–638, a mutant with deletion of PAWDETNL, accumulated in focal adhesions. Both focal adhesion numbers and area in βPIX1–638-expressing cells were greater than those in cells expressing wild-type βPIXFL. Further experiments using deletion mutants of MYO18A and βPIX showed that disruption of MYO18A–βPIX interaction not only impaired cell motility but also decreased Rac1 activity. Collectively, our data unravel the interaction regions between MYO18A and βPIX and provide evidence for the critical role of this interaction in regulating cellular localization of βPIX, Rac1 activity, and adhesion and migration in epithelial cells.  相似文献   

9.
Prostate secretory protein of 94 amino acids (PSP94) is one of the major proteins present in human seminal plasma. We had earlier reported that PSP94 has the ability to bind to human IgG. The aims of the present study were to further delineate the PSP94–IgG interaction and to understand whether this could have any significance in sperm function. Direct binding of IgG fragments to PSP94 showed maximal binding with F(ab′)2 followed by Fab, while Fc displayed least binding in ELISA. Binding kinetics of PSP94–IgG interaction using surface plasmon resonance (SPR) revealed high-affinity binding of IgG to PSP94 with a dissociation constant (KD) of 8.8 × 1011 M. PSP94–IgG interaction was found to be through the Fab domains of IgG. Real-time interaction kinetics revealed association constants for binding of IgG, Fab, and F(ab′)2 towards PSP94 to be of the same order but with altered dissociation constants. IgG and its F(ab′)2 fragment once complexed to PSP94 demonstrated negligible dissociation, while dissociation rate of Fab fragment was 6.6 × 104. In silico molecular modeling of PSP94–IgG complex identified N- and C-terminal β-strands of PSP94 to be the most plausible region involved in IgG interaction. Immunofluorescence studies revealed that IgG bound to human spermatozoa predominantly in the tail region, which could be prevented when IgG was preincubated with PSP94. This study reports for the first time that IgG forms a high-affinity complex with PSP94 through its F(ab′)2 domain and reveals the ability of PSP94 to prevent binding of IgG to spermatozoa.  相似文献   

10.
Human platelets bind on an average of 5 × 105 molecules of lentil lectin/cell with an apparent dissociation constant of 3 × 10?7 M. The lectin binds mainly to surface glycoprotein II with an apparent molecular weight of 125,000. Lentil lectin neither caused aggregation nor did it inhibit platelet aggregation by other agents. It had no influence on the binding of thrombin to platelets or on thrombin-induced clot retraction. The hypothesis that glycoprotein II mediates platelet aggregation needs reevaluation.  相似文献   

11.

Background

We have previously shown that lung collectins, surfactant protein A (SP-A) and surfactant protein D, interact with Toll-like receptor (TLR) 2, TLR4, or MD-2. Bindings of lung collectins to TLR2 and TLR4/MD-2 result in the alterations of signaling through these receptors, suggesting the immunomodulatory functions of lung collectins. Mannose binding lectin (MBL) is another collectin molecule which has structural homology to SP-A. The interaction between MBL and TLRs has not yet been determined.

Methods

We prepared recombinant MBL, and analyzed its bindings to recombinant soluble forms of TLR4 (sTLR4) and MD-2.

Results

MBL bound to sTLR4 and MD-2. The interactions were Ca2+-dependent and inhibited by mannose or monoclonal antibody against the carbohydrate-recognition domain of MBL. Treatment of sTLR4 or MD-2 by peptide N-glycosidase F significantly decreased the binding of MBL. SP-A bound to deglycosylated sTLR4, and this property did not change in chimeric molecules of SP-A/MBL in which Glu195–Phe228 or Thr174–Gly194 of SP-A were replaced with the corresponding MBL sequences.

General Significance

These results suggested that MBL binds to TLR4 and MD-2 through the carbohydrate-recognition domain, and that oligosaccharide moieties of TLR4 and MD-2 are important for recognition by MBL. Since our previous studies indicated that lung collectins bind to the peptide portions of TLRs, MBL and lung collectins interact with TLRs by different mechanisms. These direct interactions between MBL and TLR4 or MD-2 suggest that MBL may modulate cellular responses by altering signals through TLRs.  相似文献   

12.
Latex and other exudates in plants contain various proteins that are thought to play important defensive roles against herbivorous insects and pathogens. Herein, the defensive effects of phloem exudates against the Eri silkworm, Samia ricini (Saturniidae, Lepidoptera) in several cucurbitaceous plants were investigated. It was found that phloem exudates are responsible for the defensive activities of cucurbitaceous plants, such as the wax gourd Benincasa hispida and Cucumis melo, especially in B. hispida, whose leaves showed the strongest growth-inhibitory activity of all the cucurbitaceous plants tested. A 35 kDa proteinaceous growth-inhibitory factor against insects designated BPLP (B. hispida Phloem Lectin-like Protein) was next isolated and purified from the B. hispida exudate, using anion exchange and gel filtration chromatography. A very low concentration (70 μg/g) of BPLP significantly inhibited growth of S. ricini larvae. The full-length cDNA (1076 bp) encoding BPLP was cloned and its nucleotide sequence was determined. The deduced amino acid sequence of BPLP had 51% identity with a cucurbitaceous phloem lectin (phloem protein 2, PP2), and showed binding specificity to oligomers of N-acetylglucosamine. Some features of BPLP indicated that it does not have a cysteine residue and it is composed of two repeats of similar sequences, suggesting that BPLP is distinct from PP2. Recombinant BPLP, obtained by expressing the cDNA in Escherichia coli, showed both chitin-binding lectin activity and growth-inhibitory activity against S. ricini larvae. The present study thus provides experimental evidence that phloem exudates of Cucurbitaceae plants, analogous to plant latex, play defensive roles against insect herbivores, especially against chewing insects, and contain defensive substances toxic to them.  相似文献   

13.
Recent advances in genome sequencing efforts have revealed an abundance of novel putative lectins. Among these, many galectin-related proteins, characterized by many conserved residues but intriguingly lacking critical amino acids, have been found in all corners of the eukaryotic superkingdom. Here we present a structural and biochemical analysis of one representative, the galectin-related lectin CGL3 found in the inky cap mushroom Coprinopsis cinerea. This protein contains all but one conserved residues known to be involved in β-galactoside binding in galectins. A Trp residue strictly conserved among galectins is changed to an Arg in CGL3 (R81). Accordingly, the galectin-related protein is not able to bind lactose. Screening of a glycan array revealed that CGL3 displays preference for oligomers of β1-4-linked N-acetyl-glucosamines (chitooligosaccharides) and GalNAcβ1-4GlcNAc (LacdiNAc). Carbohydrate-binding affinity of this novel lectin was quantified using isothermal titration calorimetry, and its mode of chitooligosaccharide coordination not involving any aromatic amino acid residues was studied by X-ray crystallography. Structural information was used to alter the carbohydrate-binding specificity and substrate affinity of CGL3. The importance of residue R81 in determining the carbohydrate-binding specificity was demonstrated by replacing this Arg with a Trp residue (R81W). This single-amino-acid change led to a lectin that failed to bind chitooligosaccharides but gained lactose binding. Our results demonstrate that, similar to the legume lectin fold, the galectin fold represents a conserved structural framework upon which dramatically altered specificities can be grafted by few alterations in the binding site and that, in consequence, many metazoan galectin-related proteins may represent lectins with novel carbohydrate-binding specificities.  相似文献   

14.
In oilseeds whose glycerides carry common fatty acids (groundnut, sweet almond, cotton-seed, soyabean, safflower, sunflower and sesame), the constituent fatty acids of phospholipids were qualitatively but not quantitatively the same. Even when present in quantity in the triglyceride, unusual fatty acids like erucic (mustard, nasturtium, turnip), arachidic (soapnut), petroselinic (coriander) or conjugated trienes (bitter gourd, snake gourd) were totally or substantially absent in the corresponding phospholipids. In all the 14 seed oil phospholipids studied, the main fatty acids were always palmitic (17–43%), oleic (16–57%) and linoleic (10–34%).  相似文献   

15.

Background

Pathogenic bacteria specifically recognize extracellular matrix (ECM) molecules of the host (e.g. collagen, fibrinogen and fibronectin) through their surface proteins known as MSCRAMMs (Microbial Surface Components Recognizing Adhesive Matrix Molecules) and initiate colonization. On implantation, biomaterials easily get coated with these ECM molecules and the MSCRAMMs mediate bacterial adherence to biomaterials. With the rapid rise in antibiotic resistance, designing alternative strategies to reduce/eliminate bacterial colonization is absolutely essential.

Methods

The Rhusiopathiae surface protein B (RspB) is a collagen‐binding MSCRAMM of Erysipelothrix rhusiopathiae. It also binds to abiotic surfaces. The crystal structure of the collagen‐binding region of RspB (rRspB31–348) reported here revealed that RspB also binds collagen by a unique ligand binding mechanism called “Collagen Hug” which is a common theme for collagen‐binding MSCRAMMs of many Gram-positive bacteria. Here, we report the interaction studies between rRspB31–348 and silver nanoparticles using methods like gel shift assay, gel permeation chromatography and circular dichroism spectroscopy.

Results

The “Collagen Hug” mechanism was inhibited in the presence of silver nanoparticles as rRspB31–348 was unable to bind to collagen. The total loss of binding was likely because of rRspB31–348 and silver nanoparticle protein corona formation and not due to the loss of the structural integrity of rRspB31–348 on binding with nanoparticles as observed from circular dichroism experiments.

General significance

Interaction of rRspB31–348 with silver nanoparticle impaired its ligand binding mechanism. Details of this inhibition mechanism may be useful for the development of antimicrobial materials and antiadhesion drugs.  相似文献   

16.
RNA binding proteins control gene expression by the attenuation/antitermination mechanism. HutP is an RNA binding antitermination protein. It regulates the expression of hut operon when it binds with RNA by modulating the secondary structure of single-stranded hut mRNA. HutP necessitates the presence of l-histidine and divalent metal ion to bind with RNA. Herein, we report the crystal structures of ternary complex (HutP–l-histidine–Mg2+) and EDTA (0.5 M) treated ternary complex (HutP–l-histidine–Mg2+), solved at 1.9 Å and 2.5 Å resolutions, respectively, from Geobacillus thermodenitrificans. The addition of 0.5 M EDTA does not affect the overall metal-ion mediated ternary complex structure and however, the metal ions at the non-specific binding sites are chelated, as evidenced from the results of structural features.  相似文献   

17.
The 53-kDa insulin receptor substrate protein (IRSp53) organizes the actin cytoskeleton in response to stimulation of small GTPases, promoting the formation of cell protrusions such as filopodia and lamellipodia. IMD is the N-terminal 250 amino acid domain (IRSp53/MIM Homology Domain) of IRSp53 (also called I-BAR), which can bind to negatively charged lipid molecules. Overexpression of IMD induces filopodia formation in cells and purified IMD assembles finger-like protrusions in reconstituted lipid membranes. IMD was shown by several groups to bundle actin filaments, but other groups showed that it also binds to membranes. IMD binds to negatively charged lipid molecules with preference to clusters of PI(4,5)P2. Here, we performed a range of different in vitro fluorescence experiments to determine the binding properties of the IMD to phospholipids. We used different constructs of large unilamellar vesicles (LUVETs), containing neutral or negatively charged phospholipids. We found that IMD has a stronger binding interaction with negatively charged PI(4,5)P2 or PS lipids than PS/PC or neutral PC lipids. The equilibrium dissociation constant for the IMD–lipid interaction falls into the 78–170 μM range for all the lipids tested. The solvent accessibility of the fluorescence labels on the IMD during its binding to lipids is also reduced as the lipids become more negatively charged. Actin affects the IMD–lipid interaction, depending on its polymerization state. Monomeric actin partially disrupts the binding, while filamentous actin can further stabilize the IMD–lipid interaction.  相似文献   

18.
Collectins are pattern recognition molecules of the innate immune system showing binding to carbohydrate structures on microorganisms in a calcium-dependent manner. Recently, three novel collectins, collectin liver 1 (CL-L1), collectin kidney 1 (CL-K1 and CL-11), and collectin placenta 1 (CL-P1), were discovered. The roles of these three collectins remain largely unknown. Here, we present a time-resolved immunofluorometric assay for quantification of CL-L1. The concentration of CL-L1 in donor plasma (n = 210) was distributed log-normally with a median value of 3.0 μg/ml (range 1.5–5.5 μg/ml). We observed on average 30% higher concentrations of CL-L1 in plasma as compared with serum. Size analysis by gel-permeation chromatography showed CL-L1 in serum to elute as large 700–800-kDa complexes and smaller 200–300-kDa complexes. CL-L1 showed specific binding to mannose-TSK beads in a Ca2+-dependent manner. This binding could be inhibited by mannose and glucose, but not galactose, indicating that CL-L1 binds via its carbohydrate-recognition domain and has ligand specificity similar to that of mannan-binding lectin. Western blot analysis of CL-L1 showed the presence of several oligomeric forms in serum. Ontogeny studies showed CL-L1 to be present at birth at near adult levels. CL-L1 levels exhibit low variation in healthy adults over a 1-year period. During acute-phase responses, the CL-L1 levels display only minor variations. In serum, CL-L1 was found in complexes with mannan-binding lectin-associated serine proteases, suggesting a role in the lectin pathway of complement activation. The presented data establish a basis for future studies on the biological role of CL-L1.  相似文献   

19.
Abstract

The probable modes of binding of Methyl—α (and β)-D-glucopyranosides and some of their derivatives to concanavalin A have been proposed from theoretical studies. Theory predicts that βMeGlcP can bind to ConA in three different modes whereas α-MeGlcP can bind only in one mode. βMeGlcP in its most favourable mode of binding differs from α-MeGlcP in its alignment in the active-site of the lectin where it binds in a flipped or inverted orientation. Methyl substitution at the C-2 atom of the α-MeGlcP does not significantly affect the possible orientations of the sugar in the active-site of the lectin. Methyl substitution at C-3 or C-4, however, affects the allowed orientations drastically leading to the poor inhibiting power of Methyl-3-O-methyl-α-D-glucopyranoside and the inactivity of Methyl-4-O-methyl-α-D-glucopyranoside. These studies suggest that the increased activity of the α-MeGlcP over β-MeGlcP may be due to the possibility of formation of better hydrogen bonds and to hydrophobic interactions rather than to steric factors as suggested by earlier workers. These models explain the available NMR and other binding studies.  相似文献   

20.
Calponin is an actin- and calmodulin-binding protein believed to regulate the function of actin. Low-resolution studies based on proteolysis established that the recombinant calponin fragment 131–228 contained actin and calmodulin recognition sites but failed to precisely identify the actin-binding determinants. In this study, we used NMR spectroscopy to investigate the structure of this functionally important region of calponin and map its interaction with actin and calmodulin at amino-acid resolution. Our data indicates that the free calponin peptide is largely unstructured in solution, although four short amino-acid stretches corresponding to residues 140–146, 159–165, 189–195, and 199–205 display the propensity to form α-helices. The presence of four sequential transient helices probably provides the conformational malleability needed for the promiscuous nature of this region of calponin. We identified all amino acids involved in actin binding and demonstrated for the first time, to our knowledge, that the N-terminal flanking region of Lys137-Tyr144 is an integral part of the actin-binding site. We have also delineated the second actin-binding site to amino acids Thr180-Asp190. Ca2+-calmodulin binding extends beyond the previously identified minimal sequence of 153–163 and includes most amino acids within the stretch 143–165. In addition, we found that calmodulin induces chemical shift perturbations of amino acids 188–190 demonstrating for the first time, to our knowledge, an effect of Ca2+-calmodulin on this region. The spatial relationship of the actin and calmodulin contacts as well as the transient α-helical structures within the regulatory region of calponin provides a structural framework for understanding the Ca2+-dependent regulation of the actin-calponin interaction by calmodulin.  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