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.
The insect steroid hormone 20-hydroxyecdysone (20E) binds to its cognate nuclear receptor composed of the ecdysone receptor (EcR) and Ultraspiracle (USP) and triggers the main developmental transitions, in particular molting and metamorphosis. We present the crystal structure of the ligand-binding domains of EcR/USP in complex with 20E at 2.4A resolution and compare it with published structures of EcR/USP bound to ponasterone A (ponA). ponA is essentially identical to 20E but lacks the 25-OH group of 20E. The structure of 20E-bound EcR indicates that an additional hydrogen bond is formed compared with the ponA-bound receptor, yet, paradoxically, ponA has a significantly higher affinity for EcR than 20E. Theoretical studies based on docking and free energy methods lead to a rationale for understanding the difference in binding affinities between 20E and ponA. Results of the calculations indicate that the favorable contribution from the extra H-bond made by 25-OH of 20E is counterbalanced by its larger desolvation cost compared with that of ponA. The contribution of 25-OH to the binding affinity is further compared with those of 20- and 22-OH groups. Ligands that lack the 20- or 22-OH group are indeed known to bind less favorably to EcR than 20E, an effect opposite to that observed for ponA. The results indicate that their respective contributions to receptor-ligand complex stability reside mostly in their different contributions to solvation/desolvation. Together, the data demonstrate the critical role of ligand desolvation in determining binding affinity, with general implications for the binding of hormones to their cognate nuclear receptors.  相似文献   

2.
TAR DNA binding protein (TDP43), a highly conserved heterogeneous nuclear ribonucleoprotein, was found to down-regulate splicing of the exon 9 cystic fibrosis transmembrane conductance regulator (CFTR) through specific binding to a UG-rich polymorphic region upstream of the 3' splice site. Despite the emergence of new information regarding the protein's nuclear localization and splicing regulatory activity, TDP43's role in cells remains elusive. To investigate the function of human TDP43 and its homologues, we cloned and characterized the proteins from Drosophila melanogaster and Caenorhabditis elegans. The proteins from human, fly, and worm show striking similarities in their nucleic acid binding specificity. We found that residues at two different positions, which show a strong conservation among TDP43 family members, are linked to the tight recognition of the target sequence. Our three-dimensional model of TDP43 in complex with a (UG)(m) sequence predicts that these residues make amino acid side-chain to base contacts. Moreover, our results suggest that Drosophila TDP43 is comparable to human TDP43 in regulating exon splicing. On the other hand, C.elegans TDP43 has no effect on exon recognition. TDP43 from C.elegans lacks the glycine-rich domain found at the carboxy terminus of the other two homologues. Mutants of human and fly TDP43 devoid of the C-terminal domain are likewise unable to affect splicing. Our studies suggest that the glycine-rich domain is essential for splicing regulation by human and fly TDP43.  相似文献   

3.
4.
Hyperfine broadening is observable in the EPR spectrum of Brevibacterium fuscum protocatechuate 3,4-dioxygenase after lyophilization and rehydration in 17O-enriched water, demonstrating H2O ligation to the active site iron. Lack of detectable broadening in the sharp features of the spectra of three substrate complexes suggests that H2O is displaced by substrate. Water is bound in the monodentate complex with the competitive inhibitor 3-hydroxybenzoate which binds directly to the iron showing that two iron ligation sites can be occupied by nonprotein ligands. Ketonized substrate analogs which mimic a proposed transition state of the reaction cycle, 2-hydroxyisonicotinic acid N-oxide (2-OHINO) and 6-hydroxynicotinic acid N-oxide (6-OH NNO), have H2O bound in their final, bleached enzyme complexes, suggesting that these complexes are also monodentate. In contrast, a transient, initial complex of 6-OH NNO which is spectrally similar to the substrate complex, apparently does not have H2O bound. Cyanide binding occurs in two steps. The active site Fe3+ of the initial, rapidly formed, violet complex is high spin while that of the second, slowly formed, green complex is low spin; a unique state for mononuclear non-heme iron enzymes. The data suggest that the Fe-CN- and Fe-(CN-)2 complexes form sequentially. CN- binds to enzyme complexes with 2-OH INO and 6-OH NNO in one step to yield high spin Fe3+ species. In contrast, preformed substrate complexes prevent CN- binding. CN- binding eliminates the broadening due to 17O-water in the EPR spectra of both native enzyme and the enzyme-ketonized analog complexes. A model is proposed in which H2O is displaced by bidentate binding of the substrate but can potentially rebind after a subsequent substrate ketonization. The proximity of the vacatable H2O-binding site of the iron to the site of oxygen insertion suggests, however, that this site may serve to stabilize an oxygenated intermediate during the reaction cycle.  相似文献   

5.
The periplasmic iron binding protein of pathogenic Gram-negative bacteria performs an essential role in iron acquisition from transferrin and other iron sources. Structural analysis of this protein from Haemophilus influenzae identified four amino acids that ligand the bound iron: His(9), Glu(57), Tyr(195), and Tyr(196). A phosphate provides an additional ligand, and the presence of a water molecule is required to complete the octahedral geometry for stable iron binding. We report the 1.14-A resolution crystal structure of the iron-loaded form of the H. influenzae periplasmic ferric ion binding protein (FbpA) mutant H9Q. This protein was produced in the periplasm of Escherichia coli and, after purification and conversion to the apo form, was iron-loaded. H9Q is able to bind ferric iron in an open conformation. A surprising finding in the present high resolution structure is the presence of EDTA located at the previously determined anion ternary binding site, where phosphate is located in the wild type holo and apo structures. EDTA contributes four of the six coordinating ligands for iron, with two Tyr residues, 195 and 196, completing the coordination. This is the first example of a metal binding protein with a bound metal.EDTA complex. The results suggest that FbpA may have the ability to bind and transport iron bound to biological chelators, in addition to bare ferric iron.  相似文献   

6.
7.
Our investigation of the catalytic properties of Saccharomyces cerevisiae α-glucosidase (AGL) using hydroxybenzyl alcohol (HBA) isomers as transglucosylation substrates and their glucosides in hydrolytic reactions demonstrated interesting findings pertaining to the aglycon specificity of this important enzyme. AGL specificity increased from the para(p)- to the ortho(o)-HBA isomer in transglucosylation, whereas such AGL aglycon specificity was not seen in hydrolysis, thus indicating that the second step of the reaction (i.e., binding of the glucosyl acceptor) is rate-determining. To study the influence of substitution pattern on AGL kinetics, we compared AGL specificity, inferred from kinetic constants, for HBA isomers and other aglycon substrates. The demonstrated inhibitory effects of HBA isomers and their corresponding glucosides on AGL-catalyzed hydrolysis of p-nitrophenyl α-glucoside (PNPG) suggest that HBA glucosides act as competitive, whereas HBA isomers are noncompetitive, inhibitors. As such, we postulate that aromatic moieties cannot bind to an active site unless an enzyme-glucosyl complex has already formed, but they can interact with other regions of the enzyme molecule resulting in inhibition.  相似文献   

8.
A structural basis for the interaction of urea with lysozyme.   总被引:5,自引:4,他引:1       下载免费PDF全文
The effect of urea on the crystal structure of hen egg-white lysozyme has been investigated using X-ray crystallography. High resolution structures have been determined from crystals grown in the presence of 0, 0.7, 2, 3, 4, and 5 M urea and from crystals soaked in 9 M urea. All the forms are essentially isomorphous with the native type II crystals, and the derived structures exhibit excellent geometry and RMS differences from ideality in bond distances and angles. Comparison of the urea complex structures with the native enzyme (type II form, at 1.5 A resolution) indicates that the effect of urea is minimal over the concentration range studied. The mean difference in backbone conformation between the native enzyme and its urea complexes varies from 0.18 to 0.49 A. Conformational changes are limited to flexible surface loops (Thr 69-Asn 74, Ser 100-Asn 103), the active site loop (Asn 59-Cys 80), and the C-terminus (Cys 127-Leu 129). Urea molecules are bound to distinct sites on the surface of the protein. One molecule is bound to the active site cleft's C subsite, at all concentrations, in a fashion analogous to that of the N-acetyl substituent of substrate and inhibitor sugars normally bound to this site. Occupation of this subsite by urea alone does not appear to induce the conformational changes associated with inhibitor binding.  相似文献   

9.
Pyranose 2-oxidase catalyzes the oxidation of a number of carbohydrates using dioxygen; glucose, for example, is oxidized at carbon 2. The structure of pyranose 2-oxidase with the reaction product 2-keto-beta-d-glucose bound in the active center is reported in a new crystal form at 1.41 A resolution. The binding structure suggests that the alpha-anomer cannot be processed. The binding mode of the oxidized product was used to model other sugars accepted by the enzyme and to explain its specificity and catalytic rates. The reported structure at pH 6.0 shows a drastic conformational change in the loop of residues 454-461 (loop 454-461) at the active center compared to that of a closely homologous enzyme analyzed at pH 4.5 with a bound acetate inhibitor. In our structures, the loop is highly mobile and shifts to make way for the sugar to pass into the active center. Presumably, loop 454-461 functions as a gatekeeper. Apart from the wild-type enzyme, a thermostable variant was analyzed at 1.84 A resolution. In this variant, Glu542 is exchanged for a lysine. The observed stabilization could be a result of the mutated residue changing an ionic contact at a comparatively weak interface of the tetramer.  相似文献   

10.
MADS domain (for M CM1, A G, D EFA and S RF) proteins are regulatory proteins found in all major eukaryotic kingdoms. Plant MADS domain regulatory proteins have a region of moderate sequence similarity that has been designated as the K domain, and its predicted coiled-coil structure suggests a role in establishing a protein—protein interaction. In vivo studies with the Arabidopsis AGAMOUS (AG) protein have indicated that the K domain is important for AG function. Using a bait fusion protein containing the K domain and the C-terminal region of AG in a yeast two-hybrid selection, 156 clones that encode potential AG-interacting proteins were identified. These clones each encode one of four highly related MADS domain proteins: AGL2, AGL4, AGL6 and AGL9. Additional analysis showed that the K domain of AG alone was able to bind the K domains of these AGLs. This binding was further confirmed by immunoprecipitation experiments using in vitro synthesized AG and AGL K domains. These results strongly suggest that AG interacts with AGL2, AGL4, AGL6 and AGL9 in vivo. Based on these results and previous observations, it is proposed that the AG function requires interaction with at least one of these AGL proteins, and such interactions contribute to the functional specificity of the AG protein.  相似文献   

11.
The recycling of photosynthetically fixed carbon by the action of microbial glycoside hydrolases is a key biological process. The consortium of degradative enzymes involved in this process frequently display catalytic modules appended to one or more noncatalytic carbohydrate-binding modules (CBMs). CBMs play a central role in the optimization of the catalytic activity of plant cell wall hydrolases through their binding to specific plant structural polysaccharides. Despite their pivotal role in the biodegradation of plant biomass, the mechanism by which these proteins recognize their target ligands is unclear. This report describes the structure of a xylan-binding CBM (CBM15) in complex with its ligand. This module, derived from Pseudomonas cellulosa xylanase Xyn10C, binds to both soluble xylan and xylooligosaccharides. The three-dimensional crystal structure of CBM15 bound to xylopentaose has been solved by x-ray crystallography to a resolution of 1.6 A. The protein displays a similar beta-jelly roll fold to that observed in many other families of binding-modules. A groove, 20-25 A in length, on the concave surface of one of the beta-sheets presents two tryptophan residues, the faces of which are orientated at approximately 240 degrees to one another. These form-stacking interactions with the n and n+2 sugars of xylopentaose complementing the approximate 3-fold helical structure of this ligand in the binding cleft of CBM15. In four of the five observed binding subsites, the 2' and 3' hydroxyls of the bound ligand are solvent-exposed, providing an explanation for the capacity of this xylan-binding CBM to accommodate the highly decorated xylans found in the plant cell wall.  相似文献   

12.
The family 10 xylanase from Streptomyces olivaceoviridis E-86 contains a (beta/alpha)(8)-barrel as a catalytic domain, a family 13 carbohydrate binding module (CBM) as a xylan binding domain (XBD) and a Gly/Pro-rich linker between them. The crystal structure of this enzyme showed that XBD has three similar subdomains, as indicated by the presence of a triple-repeated sequence, forming a galactose binding lectin fold similar to that found in the ricin toxin B-chain. Comparison with the structure of ricin/lactose complex suggests three potential sugar binding sites in XBD. In order to understand how XBD binds to the xylan chain, we analyzed the sugar-complex structure by the soaking experiment method using the xylooligosaccharides and other sugars. In the catalytic cleft, bound sugars were observed in the xylobiose and xylotriose complex structures. In the XBD, bound sugars were identified in subdomains alpha and gamma in all of the complexes with xylose, xylobiose, xylotriose, glucose, galactose and lactose. XBD binds xylose or xylooligosaccharides at the same sugar binding sites as in the case of the ricin/lactose complex but its binding manner for xylose and xylooligosaccharides is different from the galactose binding mode in ricin, even though XBD binds galactose in the same manner as in the ricin/galactose complex. These different binding modes are utilized efficiently and differently to bind the long substrate to xylanase and ricin-type lectin. XBD can bind any xylose in the xylan backbone, whereas ricin-type lectin recognizes the terminal galactose to sandwich the large sugar chain, even though the two domains have the same family 13 CBM structure. Family 13 CBM has rather loose and broad sugar specificities and is used by some kinds of proteins to bind their target sugars. In such enzyme, XBD binds xylan, and the catalytic domain may assume a flexible position with respect to the XBD/xylan complex, inasmuch as the linker region is unstructured.  相似文献   

13.
The purified porcine recpptor for the intrinsic factor-cobalamin complex bound to concanavalin A, lentil lectin and wheat germ lectin covalently coupled to Sepharose and was eluted with the corresponding soluble sugars. In contrast, human intrinsic factor bound efficiently to concanavalin A, to some extent to lentil lectin, but only slightly to wheat germ agglutinin. The binding of IF-Cbl to the receptor was inhibited when the receptor was pre-incubated with soluble wheat germ aglutinin, with an inhibition constant estimated to be 1.9 mol/l. After transfer of the purified receptor from SDS-PAGE to Immobilon, ligand blotting of the purified receptor with iodinated lectin showed that concanavalin A and lentil lectin bound to three (75, 56 and 43 kDa) components but that wheat germ agglutinin bound only to the 75 kDa component. These results showed that the subunit of the receptor could bind to wheat germ agglutinin, resulting in an inhibition of its binding with intrinsic factor. Both binding sites of intrinsic factor and of wheat germ agglutinin could be located near to each other.  相似文献   

14.
Importin13 (Imp13) is an unusual β‐karyopherin that is able to both import and export cargoes in and out of the nucleus. In the cytoplasm, Imp13 associates with different cargoes such as Mago‐Y14 and Ubc9, and facilitates their import into the nucleus where RanGTP binding promotes the release of the cargo. In this study, we present the 2.8 Å resolution crystal structure of Imp13 in complex with the SUMO E2‐conjugating enzyme, Ubc9. The structure shows an uncommon mode of cargo–karyopherin recognition with Ubc9 binding at the N‐terminal portion of Imp13, occupying the entire RanGTP‐binding site. Comparison of the Imp13–Ubc9 complex with Imp13–Mago‐Y14 shows the remarkable plasticity of Imp13, whose conformation changes from a closed ring to an open superhelix when bound to the two different cargoes. The structure also shows that the binding mode is compatible with the sumoylated states of Ubc9. Indeed, we find that Imp13 is able to bind sumoylated Ubc9 in vitro and suppresses autosumoylation activity in the complex.  相似文献   

15.
Synercid, a new semisynthetic streptogramin-derived antibiotic containing dalfopristin and quinupristin, is used in treatment of life-threatening infections caused by glycopeptide-resistant Enterococcus faecium and other bacterial pathogens. However, dissemination of genes encoding virginiamycin acetyltransferases, enzymes that confer resistance to streptogramins, threatens to limit the medical utility of the quinupristin-dalfopristin combination. Here we present structures of virginiamycin acetyltransferase D (VatD) determined at 1.8 A resolution in the absence of ligands, at 2.8 A resolution bound to dalfopristin, and at 3.0 A resolution in the presence of acetyl-coenzyme A. Dalfopristin is bound by VatD in a similar conformation to that described previously for the streptogramin virginiamycin M1. However, specific interactions with the substrate are altered as a consequence of a conformational change in the pyrollidine ring that is propagated to adjacent constituents of the dalfopristin macrocycle. Inactivation of dalfopristin involves acetyl transfer from acetyl-coenzyme A to the sole (O-18) hydroxy group of the antibiotic that lies close to the side chain of the strictly conserved residue, His-82. Replacement of residue 82 by alanine is accompanied by a fall in specific activity of >105-fold, indicating that the imidazole moiety of His-82 is a major determinant of catalytic rate enhancement by VatD. The structure of the VatD-dalfopristin complex can be used to predict positions where further structural modification of the drug might preclude enzyme binding and thereby circumvent Synercid resistance.  相似文献   

16.
Fujihashi M  Peapus DH  Kamiya N  Nagata Y  Miki K 《Biochemistry》2003,42(38):11093-11099
Aleuria aurantia possesses a fucose-specific lectin (AAL) that is widely used as a specific probe for fucose. Fucosylated sugars often play pivotal roles in many cellular processes. We have determined the crystal structure of AAL at 2.24 A resolution in complex with only three fucose molecules in its five sugar binding sites of a six-fold beta-propeller structure. Very recently, the structure of AAL has been independently determined, showing that all the five binding sites were occupied by fucose molecules [Wimmerova, M., et al. (2003) J. Biol. Chem. 278, 27059-27067]. Stabilization of the arginine conformation bound to fucose molecules plays an essential role in generating the difference in the affinity in the five binding sites. Binding models with a couple of saccharides based on biochemical assays suggest that hydrophobic contacts also play important roles in AAL recognizing its ligand.  相似文献   

17.
The X-ray structure of mistletoe lectin I (MLI), a type-II ribosome-inactivating protein (RIP), cocrystallized with galactose is described. The model was refined at 3.0 A resolution to an R-factor of 19.9% using 21 899 reflections, with Rfree 24.0%. MLI forms a homodimer (A-B)2 in the crystal, as it does in solution at high concentration. The dimer is formed through contacts between the N-terminal domains of two B-chains involving weak polar and non-polar interactions. Consequently, the overall arrangement of sugar-binding sites in MLI differs from those in monomeric type-II RIPs: two N-terminal sugar-binding sites are 15 A apart on one side of the dimer, and two C-terminal sugar-binding sites are 87 A apart on the other side. Galactose binding is achieved by common hydrogen bonds for the two binding sites via hydroxy groups 3-OH and 4-OH and hydrophobic contact by an aromatic ring. In addition, at the N-terminal site 2-OH forms hydrogen bonds with Asp27 and Lys41, and at the C-terminal site 3-OH and 6-OH undergo water-mediated interactions and C5 has a hydrophobic contact. MLI is a galactose-specific lectin and shows little affinity for N-acetylgalactosamine. The reason for this is discussed. Structural differences among the RIPs investigated in this study (their quaternary structures, location of sugar-binding sites, and fine sugar specificities of their B-chains, which could have diverged through evolution from a two-domain protein) may affect the binding sites, and consequently the cellular transport processes and biological responses of these toxins.  相似文献   

18.
Transhydrogenase (TH) is a dimeric integral membrane enzyme in mitochondria and prokaryotes that couples proton translocation across a membrane with hydride transfer between NAD(H) and NADP(H) in soluble domains. Crystal structures of the NAD(H) binding alpha1 subunit (domain I) of Rhodospirillum rubrum TH have been determined at 1.8 A resolution in the absence of dinucleotide and at 1.9 A resolution with NADH bound. Each structure contains two domain I dimers in the asymmetric unit (AB and CD); the dimers are intimately associated and related by noncrystallographic 2-fold axes. NADH binds to subunits A and D, consistent with the half-of-the-sites reactivity of the enzyme. The conformation of NADH in subunits A and D is very similar; the nicotinamide is in the anti conformation, the A-face is exposed to solvent, and both N7 and O7 participate in hydrogen bonds. Comparison of subunits A and D to six independent copies of the subunit without bound NADH reveals multiple conformations for residues and loops surrounding the NADH site, indicating flexibility for binding and release of the substrate (product). The NADH-bound structure is also compared to the structures of R. rubrum domain I with NAD bound (PDB code 1F8G) and with NAD bound in complex with domain III of TH (PDB code 1HZZ). The NADH- vs NAD-bound domain I structures reveal conformational differences in conserved residues in the NAD(H) binding site and in dinucleotide conformation that are correlated with the net charge, i.e., oxidation state, of the nicotinamides. The comparisons illustrate how nicotinamide oxidation state can affect the domain I conformation, which is relevant to the hydride transfer step of the overall reaction.  相似文献   

19.
H Huang  M Tudor  T Su  Y Zhang  Y Hu    H Ma 《The Plant cell》1996,8(1):81-94
MADS domain proteins are members of a highly conserved family found in all eukaryotes. Genetic studies clearly indicate that many plant MADS domain proteins have different regulatory functions in flower development, yet they share a highly conserved DNA binding domain and can bind to very similar sequences. How, then, can these MADS box genes confer their specific functions? Here, we describe results from DNA binding studies of AGL1 and AGL2 (for AGAMOUS-like), two Arabidopsis MADS domain proteins that are preferentially expressed in flowers. We demonstrate that both proteins are sequence-specific DNA binding proteins and show that each binding consensus has distinct features, suggestion a mechanism for specificity. In addition, we show that the proteins with more similar amino acid sequences have more similar binding sequences. We also found that AGL2 binds to DNA in vitro as a dimer and determined the region of AGL2 that is sufficient for DNA binding and dimerization. Finally, we show that several plant MADS domain proteins can bind to DNA either as homodimers or as heterodimers, suggesting that the number of different regulators could be much greater than the number of MADS box genes.  相似文献   

20.
The binding of Fe2+ to both apo and holo mammalian ferritin has been investigated under anaerobic conditions as a function of pH. In the pH range 6.0-7.5, 8.0 +/- 0.5 Fe2+ ions bind to each apoferritin molecule, but above pH 7.5, a pH-dependent Fe2+ binding profile is observed with up to 80 Fe2+ ions binding at pH 10.0. This Fe2+ binding is reversible and is accompanied by up to two H+ being released per Fe2+ bound at pH 10.0. The Fe2+ binding to apoferritin probably occurs in the 3-fold channels. A much larger and more complex pH-dependent Fe2+ binding stoichiometry was observed for holoferritin with up to 300 Fe2+ ions binding at pH 10.0. This pH-dependent Fe2+ binding was interpreted as Fe2+ interaction at the FeOOH mineral surface with displacement of H+ from -OH or phosphate surface groups by the incoming Fe2+ ions. Mossbauer spectroscopic measurements using 57Fe-labeled Fe2+ under anaerobic conditions showed that 57Fe2+ binding to holoferritin was accompanied by electron transfer to the core, yielding 57Fe3+, presumably bound to the mineral surface. Removal of added iron by Fe2+-specific chelating agents yielded 57Fe2+, demonstrating the reversibility of this electron-transfer process. The Fe2+ bound to apo- and holoferritin is readily converted to Fe3+ by exposure to O2 and strongly retained by the respective ferritin species.  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