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.
Alanine oligomers provide a key structure for silk fibers from spider and wild silkworms.We report on structural analysis of l ‐alanyl‐l ‐alanyl‐l ‐alanyl‐l ‐alanine (Ala)4 with anti‐parallel (AP) β‐structures using X‐ray and solid‐state NMR. All of the Ala residues in the (Ala)4 are in equivalent positions, whereas for alanine trimer (Ala)3 there are two alternative locations in a unit cell as reported previously (Fawcett and Camerman, Acta Cryst., 1975, 31, 658–665). (Ala)4 with AP β‐structure is more stable than AP‐(Ala)3 due to formation of the stronger hydrogen bonds. The intermolecular structure of (Ala)4 is also different from polyalanine fiber structure, indicating that the interchain arrangement of AP β‐structure changes with increasing alanine sequencelength. Furthermore the precise 1H positions, which are usually inaccesible by X‐ray diffraction method, are determined by high resolution 1H solid state NMR combined with the chemical shift calculations by the gauge‐including projector augmented wave method. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 13–20, 2014.  相似文献   

2.
While the cis‐acyltransferase modular polyketide synthase assembly lines have largely been structurally dissected, enzymes from within the recently discovered trans‐acyltransferase polyketide synthase assembly lines are just starting to be observed crystallographically. Here we examine the ketoreductase (KR) from the first polyketide synthase module of the bacillaene nonribosomal peptide synthetase/polyketide synthase at 2.35‐Å resolution. This KR naturally reduces both α‐ and β‐keto groups and is the only KR known to do so during the biosynthesis of a polyketide. The isolated KR not only reduced an N‐acetylcysteamine‐bound β‐keto substrate to a D ‐β‐hydroxy product, but also an N‐acetylcysteamine‐bound α‐keto substrate to an L ‐α‐hydroxy product. That the substrates must enter the active site from opposite directions to generate these stereochemistries suggests that the acyl‐phosphopantetheine moiety is capable of accessing very different conformations despite being anchored to a serine residue of a docked acyl carrier protein. The features enabling stereocontrolled α‐ketoreduction may not be extensive since a KR that naturally reduces a β‐keto group within a cis‐acyltransferase polyketide synthase was identified that performs a completely stereoselective reduction of the same α‐keto substrate to generate the D ‐α‐hydroxy product. A sequence analysis of trans‐acyltransferase KRs reveals that a single residue, rather than a three‐residue motif found in cis‐acyltransferase KRs, is predictive of the orientation of the resulting β‐hydroxyl group. Proteins 2014; 82:2067–2077. © 2014 Wiley Periodicals, Inc.  相似文献   

3.
Bioinformatics studies have shown that the genomes of trypanosomatid species each encode one SCP2‐thiolase‐like protein (SLP), which is characterized by having the YDCF thiolase sequence fingerprint of the Cβ2‐Cα2 loop. SLPs are only encoded by the genomes of these parasitic protists and not by those of mammals, including human. Deletion of the Trypanosoma brucei SLP gene (TbSLP) increases the doubling time of procyclic T. brucei and causes a 5‐fold reduction of de novo sterol biosynthesis from glucose‐ and acetate‐derived acetyl‐CoA. Fluorescence analyses of EGFP‐tagged TbSLP expressed in the parasite located the TbSLP in the mitochondrion. The crystal structure of TbSLP (refined at 1.75 Å resolution) confirms that TbSLP has the canonical dimeric thiolase fold. In addition, the structures of the TbSLP‐acetoacetyl‐CoA (1.90 Å) and TbSLP‐malonyl‐CoA (2.30 Å) complexes reveal that the two oxyanion holes of the thiolase active site are preserved. TbSLP binds malonyl‐CoA tightly (Kd 90 µM), acetoacetyl‐CoA moderately (Kd 0.9 mM) and acetyl‐CoA and CoA very weakly. TbSLP possesses low malonyl‐CoA decarboxylase activity. Altogether, the data show that TbSLP is a mitochondrial enzyme involved in lipid metabolism. Proteins 2016; 84:1075–1096. © 2016 Wiley Periodicals, Inc.  相似文献   

4.
Originally annotated as the initiator of fatty acid synthesis (FAS), β‐ketoacyl‐acyl carrier protein synthase III (KAS III) is a unique component of the bacterial FAS system. Novel variants of KAS III have been identified that promote the de novo use of additional extracellular fatty acids by FAS. These KAS III variants prefer longer acyl‐groups, notably octanoyl‐CoA. Acinetobacter baumannii, a clinically important nosocomial pathogen, contains such a multifunctional KAS III (AbKAS III). To characterize the structural basis of its substrate specificity, we determined the crystal structures of AbKAS III in the presence of different substrates. The acyl‐group binding cavity of AbKAS III and co‐crystal structure of AbKAS III and octanoyl‐CoA confirmed that the cavity can accommodate acyl groups with longer alkyl chains. Interestingly, Cys264 formed a disulfide bond with residual CoA used in the crystallization, which distorted helices at the putative interface with acyl‐carrier proteins. The crystal structure of KAS III in the alternate conformation can also be utilized for designing novel antibiotics.  相似文献   

5.
ARRDC3 is one of six known human α‐arrestins, and has been implicated in the downregulation of the β2‐adrenergic receptor (β2AR). ARRDC3 consists of a two‐lobed arrestin fold and a C‐terminal tail containing two PPYX motifs. In the current model for receptor downregulation by ARRDC3, the arrestin fold portion is thought to bind the receptor, while the PPXY motifs recruit ubiquitin ligases of the NEDD4 family. Here we report the crystal structures of the N‐terminal lobe of human ARRDC3 in two conformations, at 1.73 and 2.8 Å resolution, respectively. The structures reveal a large electropositive region that is capable of binding phosphate ions of crystallization. Residues within the basic patch were shown to be important for binding to β2AR, similar to the situation with β‐arrestins. This highlights potential parallels in receptor recognition between α‐ and β‐arrestins.  相似文献   

6.
Peroxisomal β‐oxidative degradation of compounds is a common metabolic process in eukaryotes. Reported benzoyl‐coenzyme A (BA‐CoA) thioesterase activity in peroxisomes from petunia flowers suggests that, like mammals and fungi, plants contain auxiliary enzymes mediating β‐oxidation. Here we report the identification of Petunia hybrida thioesterase 1 (PhTE1), which catalyzes the hydrolysis of aromatic acyl‐CoAs to their corresponding acids in peroxisomes. PhTE1 expression is spatially, developmentally and temporally regulated and exhibits a similar pattern to known benzenoid metabolic genes. PhTE1 activity is inhibited by free coenzyme A (CoA), indicating that PhTE1 is regulated by the peroxisomal CoA pool. PhTE1 downregulation in petunia flowers led to accumulation of BA‐CoA with increased production of benzylbenzoate and phenylethylbenzoate, two compounds which rely on the presence of BA‐CoA precursor in the cytoplasm, suggesting that acyl‐CoAs can be exported from peroxisomes. Furthermore, PhTE1 downregulation resulted in increased pools of cytoplasmic phenylpropanoid pathway intermediates, volatile phenylpropenes, lignin and anthocyanins. These results indicate that PhTE1 influences (i) intraperoxisomal acyl‐CoA/CoA levels needed to carry out β‐oxidation, (ii) efflux of β‐oxidative products, acyl‐CoAs and free acids, from peroxisomes, and (iii) flux distribution within the benzenoid/phenylpropanoid metabolic network. Thus, this demonstrates that plant thioesterases play multiple auxiliary roles in peroxisomal β‐oxidative metabolism.  相似文献   

7.
The structure of trypanosomal triosephosphate isomerase (TIM)has been solved at a resolution of 2.1Å in a new crystal form grown at pH 8.8 from PEG6000. In this new crystal form (space group C2, cell dimensions 94.8 Å, 48.3 Å, 131.0 Å, 90.0°, 100.3°, 90.0°), TIM is present in a ligand-free state. The asymmetric unit consists of two TIM subunits. Each of these subunits is part of a dimer which is sitting on a crystallographic twofold axis, such that the crystal packing is formed from two TIM dimers in two distinct environments. The two constituent monomers of a given dimer are, therefore, crystallographically equivalent. In the ligand-free state of TIM in this crystal form, the two types of dimer are very similar in structure, with the flexible loops in the “Open” conformation. For one dimer (termed molecule-1), the flexible loop (loop-6) is involved in crystal contacts. Crystals of this type have been used in soaking experiments with 0.4 M ammonium sulphate (studied at 2.4 Å resolution), and with 40 μM phosphoglycolohydroxamate (studied at 2.5 Å resolution). It is found that transfer to 0.4 M ammonuum sulphate (equal to 80 times the Ki of sulphate for TIM), gives rise to significant sulphate binding at the active site of one dimer (termed molecule-2), and less significant binding at the active site of the other. In neither dimer does sulphate induce a “closed” conformation. In a mother liquor containing 40 μM phosphoglycolohydroxamate (equal to 10 times the Ki of phosphoglycolohydroxamate for TIM), an inhibitor molecule binds at the active site of only that dimer of which the flexible loop is free from crystal contacts (molecule-2). In this dimer, it induces a closed conformation. These three structures are compared and discussed with respect to the mode of binding of ligand in the active site as well as with respect to the conformational changes resulting from ligand binding. © 1993 Wiley-Liss, Inc.  相似文献   

8.
Many protein architectures exhibit evidence of internal rotational symmetry postulated to be the result of gene duplication/fusion events involving a primordial polypeptide motif. A common feature of such structures is a domain‐swapped arrangement at the interface of the N‐ and C‐termini motifs and postulated to provide cooperative interactions that promote folding and stability. De novo designed symmetric protein architectures have demonstrated an ability to accommodate circular permutation of the N‐ and C‐termini in the overall architecture; however, the folding requirement of the primordial motif is poorly understood, and tolerance to circular permutation is essentially unknown. The β‐trefoil protein fold is a threefold‐symmetric architecture where the repeating ~42‐mer “trefoil‐fold” motif assembles via a domain‐swapped arrangement. The trefoil‐fold structure in isolation exposes considerable hydrophobic area that is otherwise buried in the intact β‐trefoil trimeric assembly. The trefoil‐fold sequence is not predicted to adopt the trefoil‐fold architecture in ab initio folding studies; rather, the predicted fold is closely related to a compact “blade” motif from the β‐propeller architecture. Expression of a trefoil‐fold sequence and circular permutants shows that only the wild‐type N‐terminal motif definition yields an intact β‐trefoil trimeric assembly, while permutants yield monomers. The results elucidate the folding requirements of the primordial trefoil‐fold motif, and also suggest that this motif may sample a compact conformation that limits hydrophobic residue exposure, contains key trefoil‐fold structural features, but is more structurally homologous to a β‐propeller blade motif.  相似文献   

9.
Tryptophan 2,3‐dioxygenase (TDO), one of the two key enzymes in the kynurenine pathway, catalyzes the indole ring cleavage at the C2‐C3 bond of l ‐tryptophan. This is a rate‐limiting step in the regulation of tryptophan concentration in vivo, and is thus important in drug discovery for cancer and immune diseases. Here, we report the crystal structure of human TDO (hTDO) without the heme cofactor to 2.90 Å resolution. The overall fold and the tertiary assembly of hTDO into a tetramer, as well as the active site architecture, are well conserved and similar to the structures of known orthologues. Kinetic and mutational studies confirmed that eight residues play critical roles in l ‐tryptophan oxidation. Proteins 2014; 82:3210–3216. © 2014 Wiley Periodicals, Inc.  相似文献   

10.
Helices are important structural/recognition elements in proteins and peptides. Stability and conformational differences between helices composed of α‐ and β‐amino acids as scaffolds for mimicry of helix recognition has become a theme in medicinal chemistry. Furthermore, helices formed by β‐amino acids are experimentally more stable than those formed by α‐amino acids. This is paradoxical because the larger sizes of the hydrogen‐bonding rings required by the extra methylene groups should lead to entropic destabilization. In this study, molecular dynamics simulations using the second‐generation force field, AMOEBA (Ponder, J.W., et al., Current status of the AMOEBA polarizable force field. J Phys Chem B, 2010. 114 (8): p. 2549–64.) explored the stability and hydrogen‐bonding patterns of capped oligo‐β‐alanine, oligoalanine, and oligoglycine dodecamers in water. The MD simulations showed that oligo‐β‐alanine has strong acceptor+2 hydrogen bonds, but surprisingly did not contain a large content of 312‐helical structures, possibly due to the sparse distribution of the 312‐helical structure and other structures with acceptor+2 hydrogen bonds. On the other hand, despite its backbone flexibility, the β‐alanine dodecamer had more stable and persistent <3.0 Å hydrogen bonds. Its structure was dominated more by multicentered hydrogen bonds than either oligoglycine or oligoalanine helices. The 31 (PII) helical structure, prevalent in oligoglycine and oligoalanine, does not appear to be stable in oligo‐β‐alanine indicating its competition with other structures (stacking structure as indicated by MD analyses). These differences are among the factors that shape helical structural preferences and the relative stabilities of these three oligopeptides. Proteins 2014; 82:3043–3061. © 2014 Wiley Periodicals, Inc.  相似文献   

11.
5‐Aminolevulinate synthase (ALAS) controls the rate‐limiting step of heme biosynthesis in mammals by catalyzing the condensation of succinyl‐coenzyme A and glycine to produce 5‐aminolevulinate, coenzyme‐A (CoA), and carbon dioxide. ALAS is a member of the α‐oxoamine synthase family of pyridoxal 5′‐phosphate (PLP)‐dependent enzymes and shares high degree of structural similarity and reaction mechanism with the other members of the family. The X‐ray crystal structure of ALAS from Rhodobacter capsulatus reveals that the alkanoate component of succinyl‐CoA is coordinated by a conserved arginine and a threonine. The functions of the corresponding acyl‐CoA‐binding residues in murine erthyroid ALAS (R85 and T430) in relation to acyl‐CoA binding and substrate discrimination were examined using site‐directed mutagenesis and a series of CoA‐derivatives. The catalytic efficiency of the R85L variant with octanoyl‐CoA was 66‐fold higher than that of the wild‐type protein, supporting the proposal of this residue as key in discriminating substrate binding. Substitution of the acyl‐CoA‐binding residues with hydrophobic amino acids caused a ligand‐induced negative dichroic band at 420 nm in the CD spectra, suggesting that these residues affect substrate‐mediated changes to the PLP microenvironment. Transient kinetic analyses of the R85K variant‐catalyzed reactions confirm that this substitution decreases microscopic rates associated with formation and decay of a key reaction intermediate and show that the nature of the acyl‐CoA tail seriously affect product binding. These results show that the bifurcate interaction of the carboxylate moiety of succinyl‐CoA with R85 and T430 is an important determinant in ALAS function and may play a role in substrate specificity.  相似文献   

12.
Anamorsin is a recently identified molecule that inhibits apoptosis during hematopoiesis. It contains an N‐terminal methyltransferase‐like domain and a C‐terminal Fe‐S cluster motif. Not much is known about the function of the protein. To better understand the function of anamorsin, we have solved the crystal structure of the N‐terminal domain at 1.8 Å resolution. Although the overall structure resembles a typical S‐adenosylmethionine (SAM) dependent methyltransferase fold, it lacks one α‐helix and one β‐strand. As a result, the N‐terminal domain as well as the full‐length anamorsin did not show S‐adenosyl‐l ‐methionine (AdoMet) dependent methyltransferase activity. Structural comparisons with known AdoMet dependent methyltransferases reveals subtle differences in the SAM binding pocket that preclude the N‐terminal domain from binding to AdoMet. The N‐terminal methyltransferase‐like domain of anamorsin probably functions as a structural scaffold to inhibit methyl transfers by out‐competing other AdoMet dependant methyltransferases or acts as bait for protein–protein interactions.Proteins 2014; 82:1066–1071. © 2013 Wiley Periodicals, Inc.  相似文献   

13.
Antibiotics have, indeed, altered the course of human history as is evidenced by the increase in human life expectancy since the 1940s. Many of these natural compounds are produced by bacteria that, by necessity, must have efficient self‐resistance mechanisms. The methymycin/pikromycin producing species Streptomyces venezuelae, for example, utilizes β‐glucosylation of its macrolide products to neutralize their effects within the confines of the cell. Once released into the environment, these compounds are activated by the removal of the glucose moiety. In S. venezuelae, the enzyme responsible for removal of the sugar from the parent compound is encoded by the desR gene and referred to as DesR. It is a secreted enzyme containing 828 amino acid residues, and it is known to be a retaining glycosidase. Here, we describe the structure of the DesR/D ‐glucose complex determined to 1.4‐Å resolution. The overall architecture of the enzyme can be envisioned in terms of three regions: a catalytic core and two auxiliary domains. The catalytic core harbors the binding platform for the glucose ligand. The first auxiliary domain adopts a “PA14 fold,” whereas the second auxiliary domain contains an immunoglobulin‐like fold. Asp 273 and Glu 578 are in the proper orientation to function as the catalytic base and proton donor, respectively, required for catalysis. The overall fold of the core region places DesR into the GH3 glycoside hydrolase family of enzymes. Comparison of the DesR structure with the β‐glucosidase from Kluyveromyces marxianus shows that their PA14 domains assume remarkably different orientations.  相似文献   

14.
Human α‐amino‐β‐carboxymuconate‐ε‐semialdehyde decarboxylase determines the fate of tryptophan metabolites in the kynurenine pathway by controlling the quinolinate levels for de novo nicotinamide adenine dinucleotide biosynthesis. The unstable nature of its substrate has made gaining insight into its reaction mechanism difficult. Our electron paramagnetic resonance (EPR) spectroscopic study on the Cu‐substituted human enzyme suggests that the native substrate does not directly ligate to the metal ion. Substrate binding did not result in a change of either the hyperfine structure or the super‐hyperfine structure of the EPR spectrum. We also determined the crystal structure of the human enzyme in its native catalytically active state (at 1.99 Å resolution), a substrate analogue‐bound form (2.50 Å resolution), and a selected active site mutant form with one of the putative substrate binding residues altered (2.32 Å resolution). These structures illustrate that each asymmetric unit contains three pairs of dimers. Consistent with the EPR findings, the ligand‐bound complex structure shows that the substrate analogue does not directly coordinate to the metal ion but is bound to the active site by two arginine residues through noncovalent interactions. Proteins 2015; 83:178–187. © 2014 Wiley Periodicals, Inc.  相似文献   

15.
Plant cell cultures constitute eco‐friendly biotechnological platforms for the production of plant secondary metabolites with pharmacological activities, as well as a suitable system for extending our knowledge of secondary metabolism. Despite the high added value of taxol and the importance of taxanes as anticancer compounds, several aspects of their biosynthesis remain unknown. In this work, a genomewide expression analysis of jasmonate‐elicited Taxus baccata cell cultures by complementary DNA‐amplified fragment length polymorphism (cDNA‐AFLP) indicated a correlation between an extensive elicitor‐induced genetic reprogramming and increased taxane production in the targeted cultures. Subsequent in silico analysis allowed us to identify 15 genes with a jasmonate‐induced differential expression as putative candidates for genes encoding enzymes involved in five unknown steps of taxane biosynthesis. Among them, the TB768 gene showed a strong homology, including a very similar predicted 3D structure, with other genes previously reported to encode acyl‐CoA ligases, thus suggesting a role in the formation of the taxol lateral chain. Functional analysis confirmed that the TB768 gene encodes an acyl‐CoA ligase that localizes to the cytoplasm and is able to convert β‐phenylalanine, as well as coumaric acid, into their respective derivative CoA esters. β‐phenylalanyl‐CoA is attached to baccatin III in one of the last steps of the taxol biosynthetic pathway. The identification of this gene will contribute to the establishment of sustainable taxol production systems through metabolic engineering or synthetic biology approaches.  相似文献   

16.
Peroxisomes are thought to have played a key role in the evolution of metabolic networks of photosynthetic organisms by connecting oxidative and biosynthetic routes operating in different compartments. While the various oxidative pathways operating in the peroxisomes of higher plants are fairly well characterized, the reactions present in the primitive peroxisomes (microbodies) of algae are poorly understood. Screening of a Chlamydomonas insertional mutant library identified a strain strongly impaired in oil remobilization and defective in Cre05.g232002 (CrACX2), a gene encoding a member of the acyl‐CoA oxidase/dehydrogenase superfamily. The purified recombinant CrACX2 expressed in Escherichia coli catalyzed the oxidation of fatty acyl‐CoAs into trans‐2‐enoyl‐CoA and produced H2O2. This result demonstrated that CrACX2 is a genuine acyl‐CoA oxidase, which is responsible for the first step of the peroxisomal fatty acid (FA) β‐oxidation spiral. A fluorescent protein‐tagging study pointed to a peroxisomal location of CrACX2. The importance of peroxisomal FA β‐oxidation in algal physiology was shown by the impact of the mutation on FA turnover during day/night cycles. Moreover, under nitrogen depletion the mutant accumulated 20% more oil than the wild type, illustrating the potential of β‐oxidation mutants for algal biotechnology. This study provides experimental evidence that a plant‐type FA β‐oxidation involving H2O2‐producing acyl‐CoA oxidation activity has already evolved in the microbodies of the unicellular green alga Chlamydomonas reinhardtii.  相似文献   

17.
The resolution methods applying (?)‐(4R,5R)‐4,5‐bis(diphenylhydroxymethyl)‐2,2‐dimethyldioxolane (“TADDOL”), (?)‐(2R,3R)‐α,α,α',α'‐tetraphenyl‐1,4‐dioxaspiro[4.5]decan‐2,3‐dimethanol (“spiro‐TADDOL”), as well as the acidic and neutral Ca2+ salts of (?)‐O,O'‐dibenzoyl‐ and (?)‐O,O'‐di‐p‐toluoyl‐(2R,3R)‐tartaric acid were extended for the preparation of 1‐n‐butyl‐3‐methyl‐3‐phospholene 1‐oxide in optically active form. In one case, the intermediate diastereomeric complex could be identified by single‐crystal X‐ray analysis. The absolute P‐configuration of the enantiomers of the phospholene oxide was also determined by comparing the experimentally obtained and calculated CD spectra. Chirality 26:174–182, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

18.
19.
The structure of AF2331, a 11‐kDa orphan protein of unknown function from Archaeoglobus fulgidus, was solved by Se‐Met MAD to 2.4 Å resolution. The structure consists of an α + β fold formed by an unusual homodimer, where the two core β‐sheets are interdigitated, containing strands alternating from both subunits. The decrease in solvent‐accessible surface area upon dimerization is unusually large (3960 Å2) for a protein of its size. The percentage of the total surface area buried in the interface (41.1%) is one of the largest observed in a nonredundant set of homodimers in the PDB and is above the mean for nearly all other types of homo‐oligomers. AF2331 has no sequence homologs, and no structure similar to AF2331 could be found in the PDB using the CE, TM‐align, DALI, or SSM packages. The protein has been identified in Pfam 23.0 as the archetype of a new superfamily and is topologically dissimilar to all other proteins with the “3‐Layer (BBA) Sandwich” fold in CATH. Therefore, we propose that AF2331 forms a novel α + β fold. AF2331 contains multiple negatively charged surface clusters and is located on the same operon as the basic protein AF2330. We hypothesize that AF2331 and AF2330 may form a charge‐stabilized complex in vivo, though the role of the negatively charged surface clusters is not clear.  相似文献   

20.
《Proteins》2018,86(3):273-278
Unusual local arrangements of protein in Ramachandran space are not well represented by standard geometry tools used in either protein structure refinement using simple harmonic geometry restraints or in protein simulations using molecular mechanics force fields. In contrast, quantum chemical computations using small poly‐peptide molecular models can predict accurate geometries for any well‐defined backbone Ramachandran orientation. For conformations along transition regions—ϕ from −60 to 60°—a very good agreement with representative high‐resolution experimental X‐ray (≤1.5 Å) protein structures is obtained for both backbone C−1‐N‐Cα angle and the nonbonded O−1…C distance, while “standard geometry” leads to the “clashing” of O…C atoms and Amber FF99SB predicts distances too large by about 0.15 Å. These results confirm that quantum chemistry computations add valuable support for detailed analysis of local structural arrangements in proteins, providing improved or missing data for less understood high‐energy or unusual regions.  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