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 10 毫秒
1.
G‐protein coupled receptors (GPCRs) are transmembrane signaling molecules, with a majority of them performing important physiological roles. β2‐Adrenergic receptor (β2‐AR) is a well‐studied GPCRs that mediates natural responses to the hormones adrenaline and noradrenaline. Analysis of the ligand‐binding region of β2‐AR using the recently solved high‐resolution crystal structures revealed a number of highly conserved amino acids that might be involved in ligand binding. However, detailed structure‐function studies on some of these residues have not been performed, and their role in ligand binding remains to be elucidated. In this study, we have investigated the structural and functional role of a highly conserved residue valine 114, in hamster β2‐AR by site‐directed mutagenesis. We replaced V114 in hamster β2‐AR with a number of amino acid residues carrying different functional groups. In addition to the complementary substitutions V114I and V114L, the V114C and V114E mutants also showed significant ligand binding and agonist dependent G‐protein activation. However, the V114G, V114T, V114S, and V114W mutants failed to bind ligand in a specific manner. Molecular modeling studies were conducted to interpret these results in structural terms. We propose that the replacement of V114 influences not only the interaction of the ethanolamine side‐chains but also the aryl‐ring of the ligands tested. Results from this study show that the size and orientation of the hydrophobic residue at position V114 in β2‐AR affect binding of both agonists and antagonists, but it does not influence the receptor expression or folding.  相似文献   

2.
ATP‐citrate lyase (ACLY) catalyzes production of acetyl‐CoA and oxaloacetate from CoA and citrate using ATP. In humans, this cytoplasmic enzyme connects energy metabolism from carbohydrates to the production of lipids. In certain bacteria, ACLY is used to fix carbon in the reductive tricarboxylic acid cycle. The carboxy(C)‐terminal portion of ACLY shows sequence similarity to citrate synthase of the tricarboxylic acid cycle. To investigate the roles of residues of ACLY equivalent to active site residues of citrate synthase, these residues in ACLY from Chlorobium limicola were mutated, and the proteins were investigated using kinetics assays and biophysical techniques. To obtain the crystal structure of the C‐terminal portion of ACLY, full‐length C. limicola ACLY was cleaved, first non‐specifically with chymotrypsin and subsequently with Tobacco Etch Virus protease. Crystals of the C‐terminal portion diffracted to high resolution, providing structures that show the positions of active site residues and how ACLY tetramerizes.  相似文献   

3.
Methionine γ‐lyse (MGL) catalyzes the α, γ‐elimination of l ‐methionine and its derivatives as well as the α, β‐elimination of l ‐cysteine and its derivatives to produce α‐keto acids, volatile thiols, and ammonia. The reaction mechanism of MGL has been characterized by enzymological studies using several site‐directed mutants. The Pseudomonas putida MGL C116H mutant showed drastically reduced degradation activity toward methionine while retaining activity toward homocysteine. To understand the underlying mechanism and to discern the subtle differences between these substrates, we analyzed the crystal structures of the reaction intermediates. The complex formed between the C116H mutant and methionine demonstrated that a loop structure (Ala51–Asn64) in the adjacent subunit of the catalytic dimer cannot approach the cofactor pyridoxal 5′‐phosphate (PLP) because His116 disrupts the interaction of Asp241 with Lys240, and the liberated side chain of Lys240 causes steric hindrance with this loop. Conversely, in the complex formed between C116H mutant and homocysteine, the thiol moiety of the substrate conjugated with PLP offsets the imidazole ring of His116 via a water molecule, disrupting the interaction of His116 and Asp241 and restoring the interaction of Asp241 with Lys240. These structural data suggest that the Cys116 to His mutation renders the enzyme inactive toward the original substrate, but activity is restored when the substrate is homocysteine due to substrate‐assisted catalysis.  相似文献   

4.
FcalphaRI, the receptor specific for the Fc region of immunoglobulin A (IgA), is responsible for IgA-mediated phagocytosis, oxidative burst, and antibody-dependent cellular cytotoxicity. Using the techniques of analytical ultracentrifugation and equilibrium gel-filtration, we show that two FcalphaRI molecules bind to a single Fcalpha homodimer. Surface plasmon resonance studies confirm the 2:1 stoichiometry of binding, with equilibrium dissociation constants of 176 nM and 431 nM for the first and second binding events, respectively. The binding affinity decreases at acidic pH in a manner consistent with protonation of a single histidine residue in the binding site. A thermodynamic analysis indicates that the histidine residue does not participate in a salt-bridge in the complex; in fact, less than 10% of the free energy of binding was contributed by electrostatic interactions. The bivalent, pH-dependent interaction between FcalphaRI and IgA has important implications for cytokine-dependent phagocytosis of IgA and the FcalphaRI-mediated degradation or recycling of IgA.  相似文献   

5.
Vitronectin (Vn), a multifunctional protein of blood and extracellular matrix, interacts with complement C9. This interaction may modulate innate immunity. Details of Vn–C9 interactions are limited. Vn–C9 interactions were assessed by employing a goat homologous system and observing Vn binding to C9 in three different assays. Using recombinant fragments, C9 binding was mapped to the N‐terminus of Vn. Site directed mutagenesis was performed to alter the second arginine glycine aspartic acid (RGD) sequence (RGD‐2) of Vn. Changing R to G or D to A in RGD‐2 caused significant decrease in Vn binding to C9 whereas changing of R to G in the first RGD motif (RGD‐1) had no effect on Vn binding to C9. These results imply that the RGD‐2 of goat Vn is involved in C9 binding. In a competitive binding assay, the presence of soluble RGD peptide inhibited Vn binding to C9 whereas heparin had no effect. Vn binding to C9 was also evaluated in terms of bacterial pathogenesis. Serum dependent inhibition of Escherichia coli growth was significantly reverted when Vn or its N‐fragment were included in the assay. The C‐fragment, which did not support C9 binding, also partly nullified serum‐dependent inhibition of bacterial growth, probably through other serum component(s).  相似文献   

6.
7.
We here report the production of four biotinylated Fcγ receptor (FcγR) ectodomains and their subsequent stable capture on streptavidin‐biosensor surfaces. For receptor biotinylation, we first describe an in‐cell protocol based on the co‐transfection of two plasmids corresponding to one of the FcγR ectodomains and the BirA enzyme in mammalian cells. This strategy is compared with a standard sequential in vitro enzymatic biotinylation with respect to biotinylation level and yield. Biotinylated FcγR ectodomains that have been prepared with both strategies are then compared by analytical ultracentrifugation and surface plasmon resonance (SPR) analyses. Overall, we demonstrate that in‐cell biotinylation is an interesting alternative to standard biotinylation protocol, as it requires less purification steps while yielding higher titers. Finally, biotin‐tagged FcγRs produced with the in‐cell approach are successfully applied to the development of SPR‐based assays to evaluate the impact of the glycosylation pattern of monoclonal antibodies on their interaction with CD16a and CD64. In that endeavor, we unambiguously observe that highly galactosylated trastuzumab (TZM‐gal), non‐glycosylated trastuzumab (TZM‐NG), and reference trastuzumab are characterized by different kinetic profiles upon binding to CD16a and CD64 that had been captured at the biosensor surface via their biotin tag. More precisely, while TZM‐NG binding to CD16a was not detected, TZM‐gal formed a more stable complex with CD16a than our reference TZM. In contrast, both glycosylated TZM bound to captured CD64 in a stable and similar fashion, whereas the interaction of their non‐glycosylated form with CD64 was characterized by a higher dissociation rate. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

8.
An aldo‐keto reductase AKR5C3 from Gluconobacter oxydans (designated as Gox0644) is a useful enzyme with various substrates, including aldehydes, diacetyl, keto esters, and α‐ketocarbonyl compounds. The crystal structures of AKR5C3 in apoform in complex with NADPH and the D53A mutant (AKR5C3‐D53A) in complex with NADPH are presented herein. Structure comparison and site‐directed mutagenesis combined with biochemical kinetics analysis reveal that the conserved Asp53 in the AKR5C3 catalytic tetrad has a crucial role in securing active pocket conformation. The gain‐of‐function Asp53 to Ala mutation triggers conformational changes on the Trp30 and Trp191 side chains, improving NADPH affinity to AKR5C3, which helps increase catalytic efficiency. The highly conserved Trp30 and Trp191 residues interact with the nicotinamide moiety of NADPH and help form the NADPH‐binding pocket. The AKR5C3‐W30A and AKR5C3‐W191Y mutants show decreased activities, confirming that both residues facilitate catalysis. Residue Trp191 is in the loop structure, and the AKR5C3‐W191Y mutant does not react with benzaldehyde, which might also determine substrate recognition. Arg192, which is involved in the substrate binding, is another important residue. The introduction of R192G increases substrate‐binding affinity by improving hydrophobicity in the substrate‐binding pocket. These results not only supplement the AKRs superfamily with crystal structures but also provide useful information for understanding the catalytic properties of AKR5C3 and guiding further engineering of this enzyme.  相似文献   

9.
Drosophila melanogaster glutathione S‐transferase D3 (DmGSTD3) has a shorter amino acid sequence as compared to other GSTs known in the fruit flies. This is due to the 15 amino acid N‐terminal truncation in which normally active amino acid residue is located. The work has made use of homology modeling to visualize the arrangement of amino acid side chains in the glutathione (GSH) substrate cavity. The identified amino acids were then replaced with amino acids without functional groups in the side chains and the mutants were analyzed kinetically. Homology modeling revealed that the side chains of Y89 and Y97 were shown facing toward the substrate cavity proposing their possible role in catalyzing the conjugation. Y97A and Y89A GSH gave large changes in Km (twofold increase), Vmax (fivefold reduction), and Kcat/Km values for GSH suggesting their significant role in the conjugation reaction. The replacement at either positions has not affected the affinity of the enzyme toward 1‐chloro‐2,4‐dinitrobenzene as no significant change in values of Kmax was observed. The replacement, however, had significantly reduced the catalytic efficiency of both mutants with (Kcat/Km)GSH and (Kcat/Km)CDNB of eight‐ and twofold reduction. The recombinant DmGSTD3 has shown no activity toward 1,2‐dichloro‐4‐nitrobenzene, 2,4‐hexadienal, 2,4‐heptadienal, p‐nitrobenzyl chloride, ethacrynic acid, and sulfobromophthalein. Therefore, it was evident that DmGSTD3 has made use of distal amino acids Y97 and Y89 for GSH conjugation.  相似文献   

10.
11.
12.
Octaprenyl pyrophosphate synthase (OPPs) catalyzes consecutive condensation reactions of one allylic substrate farnesyl pyrophosphate (FPP) and five homoallylic substrate isopentenyl pyrophosphate (IPP) molecules to form a C40 long‐chain product OPP, which serves as a side chain of ubiquinone and menaquinone. OPPs belongs to the trans‐prenyltransferase class of proteins. The structures of OPPs from Escherichia coli were solved in the apo‐form as well as in complexes with IPP and a FPP thio‐analog, FsPP, at resolutions of 2.2–2.6 Å, and revealed the detailed interactions between the ligands and enzyme. At the bottom of the active‐site tunnel, M123 and M135 act in concert to form a wall which determines the final chain length. These results represent the first ligand‐bound crystal structures of a long‐chain trans‐prenyltransferase and provide new information on the mechanisms of catalysis and product chain elongation. Proteins 2015; 83:37–45. © 2014 Wiley Periodicals, Inc.  相似文献   

13.
Short‐chain dehydrogenase/reductase (SDR) is distributed in many organisms, from bacteria to humans, and has significant roles in metabolism of carbohydrates, lipids, amino acids, and other biomolecules. An important intermediate in acidic polysaccharide metabolism is 2‐keto‐3‐deoxy‐d ‐gluconate (KDG). Recently, two short and long loops in Sphingomonas KDG‐producing SDR enzymes (NADPH‐dependent A1‐R and NADH‐dependent A1‐R′) involved in alginate metabolism were shown to be crucial for NADPH or NADH coenzyme specificity. Two SDR family enzymes—KduD from Pectobacterium carotovorum (PcaKduD) and DhuD from Streptococcus pyogenes (SpyDhuD)—prefer NADH as coenzyme, although only PcaKduD can utilize both NADPH and NADH. Both enzymes reduce 2,5‐diketo‐3‐deoxy‐d ‐gluconate to produce KDG. Tertiary and quaternary structures of SpyDhuD and PcaKduD and its complex with NADH were determined at high resolution (approximately 1.6 Å) by X‐ray crystallography. Both PcaKduD and SpyDhuD consist of a three‐layered structure, α/β/α, with a coenzyme‐binding site in the Rossmann fold; similar to enzymes A1‐R and A1‐R′, both arrange the two short and long loops close to the coenzyme‐binding site. The primary structures of the two loops in PcaKduD and SpyDhuD were similar to those in A1‐R′ but not A1‐R. Charge neutrality and moderate space at the binding site of the nucleoside ribose 2′ coenzyme region were determined to be structurally crucial for dual‐coenzyme specificity in PcaKduD by structural comparison of the NADH‐ and NADPH‐specific SDR enzymes. The corresponding site in SpyDhuD was negatively charged and spatially shallow. This is the first reported study on structural determinants in SDR family KduD related to dual‐coenzyme specificity. Proteins 2016; 84:934–947. © 2016 Wiley Periodicals, Inc.  相似文献   

14.
Glutathione‐S‐transferases (GSTs) are ubiquitous detoxification enzymes that catalyse the conjugation of electrophilic substrates to glutathione. Here, we present the crystal structures of Gtt2, a GST of Saccharomyces cerevisiae, in apo and two ligand‐bound forms, at 2.23 Å, 2.20 Å and 2.10 Å, respectively. Although Gtt2 has the overall structure of a GST, the absence of the classic catalytic essential residues—tyrosine, serine and cysteine—distinguishes it from all other cytosolic GSTs of known structure. Site‐directed mutagenesis in combination with activity assays showed that instead of the classic catalytic residues, a water molecule stabilized by Ser129 and His123 acts as the deprotonator of the glutathione sulphur atom. Furthermore, only glycine and alanine are allowed at the amino‐terminus of helix‐α1 because of stereo‐hindrance. Taken together, these results show that yeast Gtt2 is a novel atypical type of cytosolic GST.  相似文献   

15.
Interference with protein–protein interactions of interfaces larger than 1500 Å2 by small drug‐like molecules is notoriously difficult, particularly if targeting homodimers. The tRNA modifying enzyme Tgt is only functionally active as a homodimer. Thus, blocking Tgt dimerization is a promising strategy for drug therapy as this protein is key to the development of Shigellosis. Our goal was to identify hot‐spot residues which, upon mutation, result in a predominantly monomeric state of Tgt. The detailed understanding of the spatial location and stability contribution of the individual interaction hot‐spot residues and the plasticity of motifs involved in the interface formation is a crucial prerequisite for the rational identification of drug‐like inhibitors addressing the respective dimerization interface. Using computational analyses, we identified hot‐spot residues that contribute particularly to dimer stability: a cluster of hydrophobic and aromatic residues as well as several salt bridges. This in silico prediction led to the identification of a promising double mutant, which was validated experimentally. Native nano‐ESI mass spectrometry showed that the dimerization of the suggested mutant is largely prevented resulting in a predominantly monomeric state. Crystal structure analysis and enzyme kinetics of the mutant variant further support the evidence for enhanced monomerization and provide first insights into the structural consequences of the dimer destabilization. Proteins 2014; 82:2713–2732. © 2014 Wiley Periodicals, Inc.  相似文献   

16.
Plants emit a variety of volatiles in response to herbivore attack, and (Z)‐3‐hexenol and its glycosides have been shown to function as defence compounds. Although the ability to incorporate and convert (Z)‐3‐hexenol to glycosides is widely conserved in plants, the enzymes responsible for the glycosylation of (Z)‐3‐hexenol remained unknown until today. In this study, uridine‐diphosphate‐dependent glycosyltransferase (UGT) candidate genes were selected by correlation analysis and their response to airborne (Z)‐3‐hexenol, which has been shown to be taken up by the tea plant. The allelic proteins UGT85A53‐1 and UGT85A53‐2 showed the highest activity towards (Z)‐3‐hexenol and are distinct from UGT85A53‐3, which displayed a similar catalytic efficiency for (Z)‐3‐hexenol and nerol. A single amino acid exchange E59D enhanced the activity towards (Z)‐3‐hexenol, whereas a L445M mutation reduced the catalytic activity towards all substrates tested. Transient overexpression of CsUGT85A53‐1 in tobacco significantly increased the level of (Z)‐3‐hexenyl glucoside. The functional characterization of CsUGT85A53 as a (Z)‐3‐hexenol UGT not only provides the foundation for the biotechnological production of (Z)‐3‐hexenyl glucoside but also delivers insights for the development of novel insect pest control strategies in tea plant and might be generally applicable to other plants.  相似文献   

17.
Cystathionine β‐lyase (CBL) catalyzes the hydrolysis of L ‐cystathionine (L ‐Cth) to produce L ‐homocysteine, pyruvate, and ammonia. A series of active‐site mutants of Escherichia coli CBL (eCBL) was constructed to investigate the roles of residues R58, R59, D116, W340, and R372 in catalysis and inhibition by aminoethoxyvinylglycine (AVG). The effects of these mutations on the kcat/K for the β‐elimination reaction range from a reduction of only 3‐fold for D116A and D116N to 6 orders of magnitude for the R372L and R372A mutants. The order of importance of these residues for the hydrolysis of L ‐Cth is: R372 >> R58 > W340 ≈ R59 > D116. Comparison of the kinetic parameters for L ‐Cth hydrolysis with those for inhibition of eCBL by AVG demonstrates that residue R58 tethers the distal carboxylate group of the substrate and confirms that residues W340 and R372 interact with the α‐carboxylate moiety. The increase in the pKa of the acidic limb and decrease in the pKa of the basic limb of the kcat/K versus pH profiles of the R58K and R58A mutants, respectively, support a role for this residue in modulating the pKa of an active‐site residue.  相似文献   

18.
There is a critical need for compounds that target cell surface integrin receptors for applications in cancer therapy and diagnosis. We used directed evolution to engineer the Ecballium elaterium trypsin inhibitor (EETI‐II), a knottin peptide from the squash family of protease inhibitors, as a new class of integrin‐binding agents. We generated yeast‐displayed libraries of EETI‐II by substituting its 6‐amino acid trypsin binding loop with 11‐amino acid loops containing the Arg‐Gly‐Asp integrin binding motif and randomized flanking residues. These libraries were screened in a high‐throughput manner by fluorescence‐activated cell sorting to identify mutants that bound to αvβ3 integrin. Select peptides were synthesized and were shown to compete for natural ligand binding to integrin receptors expressed on the surface of U87MG glioblastoma cells with half‐maximal inhibitory concentration values of 10–30 nM. Receptor specificity assays demonstrated that engineered knottin peptides bind to both αvβ3 and αvβ5 integrins with high affinity. Interestingly, we also discovered a peptide that binds with high affinity to αvβ3, αvβ5, and α5β1 integrins. This finding has important clinical implications because all three of these receptors can be coexpressed on tumors. In addition, we showed that engineered knottin peptides inhibit tumor cell adhesion to the extracellular matrix protein vitronectin, and in some cases fibronectin, depending on their integrin binding specificity. Collectively, these data validate EETI‐II as a scaffold for protein engineering, and highlight the development of unique integrin‐binding peptides with potential for translational applications in cancer. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

19.
Mutagenesis studies were carried out to examine the effects of replacement of either the nucleophile Glu-236 or the acid/base Glu-128 residue of the F/10 xylanase by a His residue. To our surprise, the affinity for the p-nitrophenyl-β- -xylobioside substrate was increased by 103-fold in the case of the mutant E128H enzyme compared with that of the wild-type F/10 xylanase. The catalytic activity of the mutant enzymes was low, despite the fact that the distance between the nucleophilic atom (an oxygen in the native xylanase and a nitrogen in the mutant) and the α-carbon was barely changed. Thus, the alteration of the acid/base functionality (Glu-128 to His mutation) provided a significantly favorable interaction within the E128H enzyme/substrate complex in the ground state, accompanying a reduction in the stabilization effect in the transition state.  相似文献   

20.
The D1‐D2 heterodimer in the reaction center core of phototrophs binds the redox plastoquinone cofactors, QA and QB, the terminal acceptors of the photosynthetic electron transfer chain in the photosystem II (PSII). This complex is the target of the herbicide atrazine, an environmental pollutant competitive inhibitor of QB binding, and consequently it represents an excellent biomediator to develop biosensors for pollutant monitoring in ecosystems. In this context, we have undertaken a study of the Chlamydomonas reinhardtii D1‐D2 proteins aimed at designing site directed mutants with increased affinity for atrazine. The three‐dimensional structure of the D1 and D2 proteins from C. reinhardtii has been homology modeled using the crystal structure of the highly homologous Thermosynechococcus elongatus proteins as templates. Mutants of D1 and D2 were then generated in silico and the atrazine binding affinity of the mutant proteins has been calculated to predict mutations able to increase PSII affinity for atrazine. The computational approach has been validated through comparison with available experimental data and production and characterization of one of the predicted mutants. The latter analyses indicated an increase of one order of magnitude of the mutant sensitivity and affinity for atrazine as compared to the control strain. Finally, D1‐D2 heterodimer mutants were designed and selected which, according to our model, increase atrazine binding affinity by up to 20 kcal/mol, representing useful starting points for the development of high affinity biosensors for atrazine.  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