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.
Bacterial translation initiation factor IF2 complexed with GTP binds to the 30S ribosomal subunit, promotes ribosomal binding of fMet‐tRNA, and favors the joining of the small and large ribosomal subunits yielding a 70S initiation complex ready to enter the translation elongation phase. Within the IF2 molecule subdomain G3, which is believed to play an important role in the IF2‐30S interaction, is positioned between the GTP‐binding G2 and the fMet‐tRNA binding C‐terminal subdomains. In this study the solution structure of subdomain G3 of Geobacillus stearothermophilus IF2 has been elucidated. G3 forms a core structure consisting of two β‐sheets with each four anti‐parallel strands, followed by a C‐terminal α‐helix. In line with its role as linker between G3 and subdomain C1, this helix has no well‐defined orientation but is endowed with a dynamic nature. The structure of the G3 core is that of a typical OB‐fold module, similar to that of the corresponding subdomain of Thermus thermophilus IF2, and to that of other known RNA‐binding modules such as IF2‐C2, IF1 and subdomains II of elongation factors EF‐Tu and EF‐G. Structural comparisons have resulted in a model that describes the interaction between IF2‐G3 and the 30S ribosomal subunit.  相似文献   

2.
The plant actin depolymerizing factor (ADF) binds to both monomeric and filamentous actin, and is directly involved in the depolymerization of actin filaments. To better understand the actin binding sites of the Arabidopsis thaliana L. AtADF1, we generated mutants of AtADF1 and investigated their functions in vitro and in vivo. Analysis of mutants harboring amino acid substitutions revealed that charged residues (Arg98 and Lys100) located at the α‐helix 3 and forming an actin binding site together with the N‐terminus are essential for both G‐ and F‐actin binding. The basic residues on the β‐strand 5 (K82/A) and the α‐helix 4 (R135/A, R137/A) form another actin binding site that is important for F‐actin binding. Using transient expression of CFP‐tagged AtADF1 mutant proteins in onion (Allium cepa) peel epidermal cells and transgenic Arabidopsis thaliana L. plants overexpressing these mutants, we analyzed how these mutant proteins regulate actin organization and affect seedling growth. Our results show that the ADF mutants with a lower affinity for actin filament binding can still be functional, unless the affinity for actin monomers is also affected. The G‐actin binding activity of the ADF plays an essential role in actin binding, depolymerization of actin polymers, and therefore in the control of actin organization.  相似文献   

3.
4.
Encephalitozoon cuniculi is a unicellular, obligate intracellular eukaryotic parasite in the Microsporidia family and one of the agents responsible for microsporidosis infections in humans. Like most Microsporidia, the genome of E. cuniculi is markedly reduced and the organism contains mitochondria‐like organelles called mitosomes instead of mitochondria. Here we report the solution NMR structure for a protein physically associated with mitosome‐like organelles in E. cuniculi, the 128‐residue, adrenodoxin‐like protein Ec‐Adx (UniProt ID Q8SV19) in the [2Fe‐2S] ferredoxin superfamily. Oxidized Ec‐Adx contains a mixed four‐strand β‐sheet, β2‐β1‐β4‐β3 (↓↑↑↓), loosely encircled by three α‐helices and two 310‐helices. This fold is similar to the structure observed in other adrenodoxin and adrenodoxin‐like proteins except for the absence of a fifth anti‐parallel β‐strand next to β3 and the position of α3. Cross peaks are missing or cannot be unambiguously assigned for 20 amide resonances in the 1H‐15N HSQC spectrum of Ec‐Adx. These missing residues are clustered primarily in two regions, G48‐V61 and L94‐L98, containing the four cysteine residues predicted to ligate the paramagnetic [2Fe‐2S] cluster. Missing amide resonances in 1H‐15N HSQC spectra are detrimental to NMR‐based solution structure calculations because 1H‐1H NOE restraints are absent (glass half‐empty) and this may account for the absent β‐strand (β5) and the position of α3 in oxidized Ec‐Adx. On the other hand, the missing amide resonances unambiguously identify the presence, and immediate environment, of the paramagnetic [2Fe‐2S] cluster in oxidized Ec‐Adx (glass half‐full).  相似文献   

5.
Gas vesicles are proteinaceous, gas‐filled nanostructures produced by some bacteria and archaea. The hydrophobic major structural protein GvpA forms the ribbed gas vesicle wall. An in‐silico 3D‐model of GvpA of the predicted coil‐α1‐β1‐β2‐α2‐coil structure is available and implies that the two β‐chains constitute the hydrophobic interior surface of the gas vesicle wall. To test the importance of individual amino acids in GvpA we performed 85 single substitutions and analyzed these variants in Haloferax volcanii ΔA + Amut transformants for their ability to form gas vesicles (Vac+ phenotype). In most cases, an alanine substitution of a non‐polar residue did not abolish gas vesicle formation, but the replacement of single non‐polar by charged residues in β1 or β2 resulted in Vac transformants. A replacement of residues near the β‐turn altered the spindle‐shape to a cylindrical morphology of the gas vesicles. Vac transformants were also obtained with alanine substitutions of charged residues of helix α1 suggesting that these amino acids form salt‐bridges with another GvpA monomer. In helix α2, only the alanine substitution of His53 or Tyr54, led to Vac transformants, whereas most other substitutions had no effect. We discuss our results in respect to the GvpA structure and data available from solid‐state NMR.  相似文献   

6.
The sensor histidine kinases of two‐component signal‐transduction systems (TCSs) are essential for bacteria to adapt to variable environmental conditions. The two‐component regulatory system BaeS/R increases multidrug and metal resistance in Salmonella and Escherichia coli. In this study, we report the X‐ray structure of the periplasmic sensor domain of BaeS from Serratia marcescens FS14. The BaeS sensor domain (34–160) adopts a mixed α/β‐fold containing a central four‐stranded antiparallel β‐sheet flanked by a long N‐terminal α‐helix and additional loops and a short C‐terminal α‐helix on each side. Structural comparisons revealed that it belongs to the PDC family with a remarkable difference in the orientation of the helix α2. In the BaeS sensor domain, this helix is situated perpendicular to the long helix α1 and holds helix α1 in the middle with the beta sheet, whereas in other PDC domains, helix α2 is parallel to helix α1. Because the helices α1 and α2 is involved in the dimeric interface, this difference implies that BaeS uses a different dimeric interface compared with other PDC domains. Proteins 2017; 85:1784–1790. © 2017 Wiley Periodicals, Inc.  相似文献   

7.
8.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Plant actin depolymerizing factor (ADF) binds to both monomeric and filamentous actin, and plays a key role in the organization of the actin cytoskeleton. In this issue, Dong et al. (pp. 250–261) demonstrate that charged residues Arg98 and Lys100 of ADF1 are essential for both G‐ and F‐actin binding, and that basic residues on β‐strand 5 (K82/A) and α‐helix 4 (R135/A, R137/A) form another actin binding site for F‐actin.  相似文献   

10.
A 34‐residue α/β peptide [IG(28–61)], derived from the C‐terminal part of the B3 domain of the immunoglobulin binding protein G from Streptoccocus, was studied using CD and NMR spectroscopy at various temperatures and by differential scanning calorimetry. It was found that the C‐terminal part (a 16‐residue‐long fragment) of this peptide, which corresponds to the sequence of the β‐hairpin in the native structure, forms structure similar to the β‐hairpin only at T = 313 K, and the structure is stabilized by non‐native long‐range hydrophobic interactions (Val47–Val59). On the other hand, the N‐terminal part of IG(28–61), which corresponds to the middle α‐helix in the native structure, is unstructured at low temperature (283 K) and forms an α‐helix‐like structure at 305 K, and only one helical turn is observed at 313 K. At all temperatures at which NMR experiments were performed (283, 305, and 313 K), we do not observe any long‐range connectivities which would have supported packing between the C‐terminal (β‐hairpin) and the N‐terminal (α‐helix) parts of the sequence. Such interactions are absent, in contrast to the folding pathway of the B domain of protein G, proposed recently by Kmiecik and Kolinski (Biophys J 2008, 94, 726–736), based on Monte‐Carlo dynamics studies. Alternative folding mechanisms are proposed and discussed. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 469–480, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

11.
Ups1 forms a complex with Mdm35 and is critical for the transport of phosphatidic acid (PA) from the mitochondrial outer membrane to the inner membrane. We report the crystal structure of the Ups1‐Mdm35‐PA complex and the functional characterization of Ups1‐Mdm35 in PA binding and transfer. Ups1 features a barrel‐like structure consisting of an antiparallel β‐sheet and three α‐helices. Mdm35 adopts a three‐helical clamp‐like structure to wrap around Ups1 to form a stable complex. The β‐sheet and α‐helices of Ups1 form a long tunnel‐like pocket to accommodate the substrate PA, and a short helix α2 acts as a lid to cover the pocket. The hydrophobic residues lining the pocket and helix α2 are critical for PA binding and transfer. In addition, a hydrophilic patch on the surface of Ups1 near the PA phosphate‐binding site also plays an important role in the function of Ups1‐Mdm35. Our study reveals the molecular basis of the function of Ups1‐Mdm35 and sheds new light on the mechanism of intramitochondrial phospholipid transport by the MSF1/PRELI family proteins.  相似文献   

12.
In many prokaryotic organisms, chromosomal loci known as clustered regularly interspaced short palindromic repeats (CRISPRs) and CRISPR‐associated (CAS) genes comprise an acquired immune defense system against invading phages and plasmids. Although many different Cas protein families have been identified, the exact biochemical functions of most of their constituents remain to be determined. In this study, we report the crystal structure of PF1127, a Cas protein of Pyrococcus furiosus DSM 3638 that is composed of 480 amino acids and belongs to the Csx1 family. The C‐terminal domain of PF1127 has a unique β‐hairpin structure that protrudes out of an α‐helix and contains several positively charged residues. We demonstrate that PF1127 binds double‐stranded DNA and RNA and that this activity requires an intact β‐hairpin and involve the homodimerization of the protein. In contrast, another Csx1 protein from Sulfolobus solfataricus P2 that is composed of 377 amino acids does not have the β‐hairpin structure and exhibits no DNA‐binding properties under the same experimental conditions. Notably, the C‐terminal domain of these two Csx1 proteins is greatly diversified, in contrast to the conserved N‐terminal domain, which appears to play a common role in the homodimerization of the protein. Thus, although P. furiosus Csx1 is identified as a nucleic acid‐binding protein, other Csx1 proteins are predicted to exhibit different individual biochemical activities. Proteins 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

13.
HP1454 is a protein of 303 amino acids found in the extracellular milieu of Helicobacter pylori. The protein structure, crystallized in the orthorhombic C2221 space group with one molecule per asymmetric unit, has been determined using the single‐wavelength anomalous dispersion method. HP1454 exhibits an elongated bent shape, composed of three distinct domains. Each domain possesses a fold already present in other structures: Domain I contains a three‐strand antiparallel β‐barrel flanked by a long α‐helix, Domain II is an anti‐parallel three‐helix bundle, and Domain III a β‐sheet flanked by two α‐helices. The overall assembly of the protein does not bear any similarity with known structures. Proteins 2014; 82:2868–2873. © 2014 Wiley Periodicals, Inc.  相似文献   

14.
Plants use a highly evolved immune system to exhibit defense response against microbial infections. The plant TIR domain, together with the nucleotide‐binding (NB) domain and/or a LRR region, forms a type of molecule, named resistance (R) proteins, that interact with microbial effector proteins and elicit hypersensitive responses against infection. Here, we report the first crystal structure of a plant TIR domain from Arabidopsis thaliana (AtTIR) solved at a resolution of 2.0 Å. The structure consists of five β‐strands forming a parallel β‐sheet at the core of the protein. The β‐strands are connected by a series of α‐helices and the overall fold mimics closely that of other mammalian and bacterial TIR domains. However, the region of the αD‐helix reveals significant differences when compared with other TIR structures, especially the αD3‐helix that corresponds to an insertion only present in plant TIR domains. Available mutagenesis data suggest that several conserved and exposed residues in this region are involved in the plant TIR signaling function.  相似文献   

15.
ES Elgin  N Sökmen  FC Peterson  BF Volkman  C Dağ  AL Haas 《Proteins》2012,80(10):2482-2487
The covalent attachment of ubiquitin (Ub) and ubiquitin‐like (Ubl) proteins to various eukaryotic targets plays critical roles in regulating numerous cellular processes. E1‐activating enzymes are critical, because they catalyze activation of their cognate Ub/Ubl protein and are responsible for its transfer to the correct E2‐conjugating enzyme(s). The activating enzyme for neural‐precursor‐cell‐expressed developmentally downregulated 8 (NEDD8) is a heterodimer composed of APPBP1 and Uba3 subunits. The carboxyl terminal ubiquitin‐like β‐grasp domain of human Uba3 (Uba3‐βGD) has been suggested as a key E2‐binding site defining E2 specificity. In crystal structures of free E1 and the NEDD8‐E1 complex, the E2‐binding surface on the domain was missing from the electron density. However, when complexed with various E2s, this missing segment adopts a kinked α‐helix. Here, we demonstrate that Uba3‐βGD is an independently folded domain in solution and that residues involved in E2 binding are absent from the NMR spectrum, indicating that the E2‐binding surface on Uba3‐βGD interconverts between multiple conformations, analogous to a similar conformational transition observed in the E2‐binding surface of SUMO E1. These results suggest that access to multiple conformational substates is an important feature of the E1–E2 interaction. Proteins 2012;. © 2012 Wiley Periodicals, Inc.  相似文献   

16.
Epithelial–mesenchymal transition (EMT), via activation of Wnt signaling, is prevailing in embryogenesis, but postnatally it only occurs in pathological processes, such as in tissue fibrosis and tumor metastasis. Our prior studies led us to speculate that EMT might be involved in the loss of limbal epithelial stem cells in explant cultures. To examine this hypothesis, we successfully grew murine corneal/limbal epithelial progenitors by prolonging the culture time and by seeding at a low density in a serum‐free medium. Single cell‐derived clonal growth was accompanied by a gradient of Wnt signaling activity, from the center to the periphery, marked by a centrifugal loss of E‐cadherin and β‐catenin from intercellular junctions, coupled with nuclear translocation of β‐catenin and LEF‐1. Large‐colony‐forming efficiency at central location of colony was higher than peripheral location. Importantly, there was also progressive centrifugal differentiation, with positive K14 keratin expression and the loss of p63 and PCNA nuclear staining, and irreversible EMT, evidenced by cytoplasmic expression of α‐SMA and nuclear localization of S100A4; and by nuclear translocation of Smad4. Furthermore, cytoplasmic expression of α‐SMA was promoted by high‐density cultures and their conditioned media, which contained cell density‐dependent levels of TGF‐β1, TGF‐β2, GM‐CSF, and IL‐1α. Exogenous TGF‐β1 induced α‐SMA positive cells in a low‐density culture, while TGF‐β1 neutralizing antibody partially inhibited α‐SMA expression in a high‐density culture. Collectively, these results indicate that irreversible EMT emerges in the periphery of clonal expansion where differentiation and senescence of murine corneal/limbal epithelial progenitors occurs as a result of Smad‐mediated TGF‐β‐signaling. J. Cell. Physiol. 228: 225–234, 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

17.
Microtubule (MT) binding accelerates the rate of ATP hydrolysis in kinesin. To understand the underlying mechanism, using charged‐to‐alanine mutational analysis, we identified two independent sites in tubulin, which are critical for kinesin motility, namely, a cluster of negatively charged residues spanning the helix 11–12 (H11–12) loop and H12 of α‐tubulin, and the negatively charged residues in H12 of β‐tubulin. Mutation in the α‐tubulin‐binding site results in a deceleration of ATP hydrolysis (kcat), whereas mutation in the β‐tubulin‐binding site lowers the affinity for MTs (K0.5MT). The residue E415 in α‐tubulin seems to be important for coupling MT binding and ATPase activation, because the mutation at this site results in a drastic reduction in the overall rate of ATP hydrolysis, largely due to a deceleration in the reaction of ADP release. Our results suggest that kinesin binding at a region containing α‐E415 could transmit a signal to the kinesin nucleotide pocket, triggering its conformational change and leading to the release of ADP.  相似文献   

18.
Translation initiation factor IF3 is required for peptide chain initiation in Escherichia coli. IF3 binds directly to 30S ribosomal subunits ensuring a constant supply of free 30S subunits for initiation complex formation, participates in the kinetic selection of the correct initiator region of mRNA, and destabilizes initiation complexes containing noninitiator tRNAs. The roles that tyrosine 107 and lysine 110 play in IF3 function were examined by site-directed mutagenesis. Tyrosine 107 was changed to either phenylalanine (Y107F) or leucine (Y107L), and lysine 110 was converted to either arginine (K110R) or leucine (K110L). These single amino acid changes resulted in a reduced affinity of IF3 for 30S subunits. Association equilibrium constants (M-1) for 30S subunit binding were as follows: wild-type, 7.8 x 10(7); Y107F, 4.1 x 10(7); Y107L, 1 x 10(7); K110R, 5.1 x 10(6); K110L, < 1 x 10(2). The mutant IF3s were similarly impaired in their abilities to specifically select initiation complexes containing tRNA(fMet). Toeprint analysis indicated that 5-fold more Y107L or K110R protein was required for proper initiator tRNA selection. K110L protein was unable to mediate this selection even at concentrations up to 10-fold higher than wild type. The results indicate that tyrosine 107 and lysine 110 are critical components of the ribosome binding domain of IF3 and, furthermore, that dissociation of complexes containing noninitiator tRNAs requires prior binding of IF3 to the ribosomes.  相似文献   

19.
Microbial β‐1,4‐galactanases are glycoside hydrolases belonging to family 53, which degrade galactan and arabinogalactan side chains in the hairy regions of pectin, a major plant cell wall component. They belong to the larger clan GH‐A of glycoside hydrolases, which cover many different poly‐ and oligosaccharidase specificities. Crystallographic complexes of Bacillus licheniformis β‐1,4‐galactanase and its inactive nucleophile mutant have been obtained with methyl‐β(1→4)‐galactotetraoside, providing, for the first time, information on substrate binding to the aglycone side of the β‐1,4‐galactanase substrate binding groove. Using the experimentally determined subsites as a starting point, a β(1→4)‐galactononaose was built into the structure and subjected to molecular dynamics simulations giving further insight into the residues involved in the binding of the polysaccharide from subsite ?4 to +5. In particular, this analysis newly identified a conserved β‐turn, which contributes to subsites ?2 to +3. This β‐turn is unique to family 53 β‐1,4‐galactanases among all clan GH‐A families that have been structurally characterized and thus might be a structural signature for endo‐β‐1,4‐galactanase specificity. Proteins 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

20.
Cyclic nucleotide‐sensitive ion channels, known as HCN and CNG channels, are crucial in neuronal excitability and signal transduction of sensory cells. HCN and CNG channels are activated by binding of cyclic nucleotides to their intracellular cyclic nucleotide‐binding domain (CNBD). However, the mechanism by which the binding of cyclic nucleotides opens these channels is not well understood. Here, we report the solution structure of the isolated CNBD of a cyclic nucleotide‐sensitive K+ channel from Mesorhizobium loti. The protein consists of a wide anti‐parallel β‐roll topped by a helical bundle comprising five α‐helices and a short 310‐helix. In contrast to the dimeric arrangement (‘dimer‐of‐dimers’) in the crystal structure, the solution structure clearly shows a monomeric fold. The monomeric structure of the CNBD supports the hypothesis that the CNBDs transmit the binding signal to the channel pore independently of each other.  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