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.
α‐Aminoadipate aminotransferase (AAA‐AT) catalyzes the amination of 2‐oxoadipate to α‐aminoadipate in the fourth step of the α‐aminoadipate pathway of lysine biosynthesis in fungi. The aromatic aminotransferase Aro8 has recently been identified as an AAA‐AT in Saccharomyces cerevisiae. This enzyme displays broad substrate selectivity, utilizing several amino acids and 2‐oxo acids as substrates. Here we report the 1.91Å resolution crystal structure of Aro8 and compare it to AAA‐AT LysN from Thermus thermophilus and human kynurenine aminotransferase II. Inspection of the active site of Aro8 reveals asymmetric cofactor binding with lysine‐pyridoxal‐5‐phosphate bound within the active site of one subunit in the Aro8 homodimer and pyridoxamine phosphate and a HEPES molecule bound to the other subunit. The HEPES buffer molecule binds within the substrate‐binding site of Aro8, yielding insights into the mechanism by which it recognizes multiple substrates and how this recognition differs from other AAA‐AT/kynurenine aminotransferases.  相似文献   

2.
Aims: The 2‐deoxystreptamine‐containing aminoglycoside antibiotics (AGAs) constitute the largest subgroup of the aminoglycosides. Neomycin (NEO) and lividomycin (LIV) are both representatives of the pseudo‐tetrasaccharide group among the NEO‐type AGAs. While NEO contains a 6′‐NH2 group, the 6′‐position remains unmodified in LIV. The aim of the study was to characterize the substrate specificities of the enzymes involved in the C‐6′‐ and C‐6?‐modification in order to explain the different amination patterns. Methods and Results: We overproduced and purified the enzymes NeoQ (bifunctional 6′‐ and 6?‐oxidoreductase) and NeoB (bifunctional 6′‐ and‐6?‐aminotransferase), which had been analysed before ( Huang et al. 2007 ), and compared the enzymatic properties with the corresponding enzymes LivQ (postulated 6?‐oxidoreductase, 72% identity to NeoQ) and LivB (postulated 6?‐aminotransferase, 71% identity to NeoB) from the LIV pathway. By applying a newly established photometric assay, we proved that LivQ oxidized only pseudotetrasaccharidic substrates at the 6?‐position. In contrast, NeoQ accepted also the pseudodisaccharidic paromamine as a substrate and oxidized the 6′‐ and 6?‐positions on two different precursors of NEO. The aminotransferases LivB and NeoB both transfer NH2 groups to the 6′‐position in the precursor 6′‐oxo‐paromamine and to the 6?‐position of 6?‐oxo‐neomycin C. Conclusions: The difference in the modification pattern of NEO and LIV at their 6′‐positions is based only on the difference in the substrate specificities of the oxidoreductases LivQ and NeoQ, respectively. The aminotransferases LivB and NeoB share identical biochemical properties, and both are capable to transaminate the 6′ and also the 6?‐position of the tested AGAs. Significance and Impact of the Study: Our data provide information to understand the structural variations in aminoglycosides and may be helpful to interpret variations in other natural product bisoynthesis pathways.  相似文献   

3.
N‐formylated sugars have been observed on the O‐antigens of such pathogenic Gram‐negative bacteria as Campylobacter jejuni and Francisella tularensis. Until recently, however, little was known regarding the overall molecular architectures of the N‐formyltransferases that are required for the biosynthesis of these unusual sugars. Here we demonstrate that the protein encoded by the wbtj gene from F. tularensis is an N‐formyltransferase that functions on dTDP‐4‐amino‐4,6‐dideoxy‐d ‐glucose as its substrate. The enzyme, hereafter referred to as WbtJ, demonstrates a strict requirement for N10‐formyltetrahydrofolate as its carbon source. In addition to the kinetic analysis, the three‐dimensional structure of the enzyme was solved in the presence of dTDP‐sugar ligands to a nominal resolution of 2.1 Å. Each subunit of the dimeric enzyme is dominated by a “core” domain defined by Met 1 to Ser 185. This core motif harbors the active site residues. Following the core domain, the last 56 residues fold into two α‐helices and a β‐hairpin motif. The hairpin motif is responsible primarily for the subunit:subunit interface, which is characterized by a rather hydrophobic pocket. From the study presented here, it is now known that WbtJ functions on C‐4′ amino sugars. Another enzyme recently investigated in the laboratory, WlaRD, formylates only C‐3′ amino sugars. Strikingly, the quaternary structures of WbtJ and WlaRD are remarkably different. In addition, there are several significant variations in the side chains that line their active site pockets, which may be important for substrate specificity. Details concerning the kinetic and structural properties of WbtJ are presented.  相似文献   

4.
5.
N,N'‐diacetylbacillosamine is a novel sugar that plays a key role in bacterial glycosylation. Three enzymes are required for its biosynthesis in Campylobacter jejuni starting from UDP‐GlcNAc. The focus of this investigation, PglE, catalyzes the second step in the pathway. It is a PLP‐dependent aminotransferase that converts UDP‐2‐acetamido‐4‐keto‐2,4,6‐trideoxy‐d ‐glucose to UDP‐2‐acetamido‐4‐amino‐2,4,6‐trideoxy‐d ‐glucose. For this investigation, the structure of PglE in complex with an external aldimine was determined to a nominal resolution of 2.0 Å. A comparison of its structure with those of other sugar aminotransferases reveals a remarkable difference in the manner by which PglE accommodates its nucleotide‐linked sugar substrate.  相似文献   

6.
Pantoea ananatis is a Gram‐negative bacterium first recognized in 1928 as the causative agent of pineapple rot in the Philippines. Since then various strains of the organism have been implicated in the devastation of agriculturally important crops. Some strains, however, have been shown to function as non‐pathogenic plant growth promoting organisms. To date, the factors that determine pathogenicity or lack thereof between the various strains are not well understood. All P. ananatis strains contain lipopolysaccharides, which differ with respect to the identities of their associated sugars. Given our research interest on the presence of the unusual sugar, 4‐formamido‐4,6‐dideoxy‐d ‐glucose, found on the lipopolysaccharides of Campylobacter jejuni and Francisella tularensis, we were curious as to whether other bacteria have the appropriate biosynthetic machinery to produce these unique carbohydrates. Four enzymes are typically required for their biosynthesis: a thymidylyltransferase, a 4,6‐dehydratase, an aminotransferase, and an N‐formyltransferase. Here, we report that the gene SAMN03097714_1080 from the P. ananatis strain NFR11 does, indeed, encode for an N‐formyltransferase, hereafter referred to as PA1080c. Our kinetic analysis demonstrates that PA1080c displays classical Michaelis–Menten kinetics with dTDP‐4‐amino‐4,6‐dideoxy‐d ‐glucose as the substrate and N10‐formyltetrahydrofolate as the carbon source. In addition, the X‐ray structure of PA1080c, determined to 1.7 Å resolution, shows that the enzyme adopts the molecular architecture observed for other sugar N‐formyltransferases. Analysis of the P. ananatis NFR11 genome suggests that the three other enzymes necessary for N‐formylated sugar biosynthesis are also present. Intriguingly, those strains of P. ananatis that are non‐pathogenic apparently do not contain these genes.  相似文献   

7.
Cook PD  Carney AE  Holden HM 《Biochemistry》2008,47(40):10685-10693
Perosamine (4-amino-4,6-dideoxy- d-mannose), or its N-acetylated form, is one of several dideoxy sugars found in the O-antigens of such infamous Gram-negative bacteria as Vibrio cholerae O1 and Escherichia coli O157:H7. It is added to the bacterial O-antigen via a nucleotide-linked version, namely GDP-perosamine. Three enzymes are required for the biosynthesis of GDP-perosamine starting from mannose 1-phosphate. The focus of this investigation is GDP-perosamine synthase from Caulobacter crescentus, which catalyzes the final step in GDP-perosamine synthesis, the conversion of GDP-4-keto-6-deoxymannose to GDP-perosamine. The enzyme is PLP-dependent and belongs to the aspartate aminotransferase superfamily. It contains the typically conserved active site lysine residue, which forms a Schiff base with the PLP cofactor. Two crystal structures were determined for this investigation: a site-directed mutant protein (K186A) complexed with GDP-perosamine and the wild-type enzyme complexed with an unnatural ligand, GDP-3-deoxyperosamine. These structures, determined to 1.6 and 1.7 A resolution, respectively, revealed the manner in which products, and presumably substrates, are accommodated within the active site pocket of GDP-perosamine synthase. Additional kinetic analyses using both the natural and unnatural substrates revealed that the K m for the unnatural substrate was unperturbed relative to that of the natural substrate, but the k cat was lowered by a factor of approximately 200. Taken together, these studies shed light on why GDP-perosamine synthase functions as an aminotransferase whereas another very similar PLP-dependent enzyme, GDP-4-keto-6-deoxy- d-mannose 3-dehydratase or ColD, catalyzes a dehydration reaction using the same substrate.  相似文献   

8.
The aminotransferase (BtrR), which is involved in the biosynthesis of butirosin, a 2-deoxystreptamine (2-DOS)-containing aminoglycoside antibiotic produced by Bacillus circulans, catalyses the pyridoxal phosphate (PLP)-dependent transamination reaction both of 2-deoxy-scyllo-inosose to 2-deoxy-scyllo-inosamine and of amino-dideoxy-scyllo-inosose to 2-DOS. The high-resolution crystal structures of the PLP- and PMP-bound forms of BtrR aminotransferase from B. circulans were solved at resolutions of 2.1 A and 1.7 A with R(factor)/R(free) values of 17.4/20.6 and 19.9/21.9, respectively. BtrR has a fold characteristic of the aspartate aminotransferase family, and sequence and structure analysis categorises it as a member of SMAT (secondary metabolite aminotransferases) subfamily. It exists as a homodimer with two active sites per dimer. The active site of the BtrR protomer is located in a cleft between an alpha helical N-terminus, a central alphabetaalpha sandwich domain and an alphabeta C-terminal domain. The structures of the PLP- and PMP-bound enzymes are very similar; however BtrR-PMP lacks the covalent bond to Lys192. Furthermore, the two forms differ in the side-chain conformations of Trp92, Asp163, and Tyr342 that are likely to be important in substrate selectivity and substrate binding. This is the first three-dimensional structure of an enzyme from the butirosin biosynthesis gene cluster.  相似文献   

9.
AT2433 from Actinomadura melliaura is an indolocarbazole antitumor antibiotic structurally distinguished by its unique aminodideoxypentose‐containing disaccharide moiety. The corresponding sugar nucleotide‐based biosynthetic pathway for this unusual sugar derives from comparative genomics where AtmS13 has been suggested as the contributing sugar aminotransferase (SAT). Determination of the AtmS13 X‐ray structure at 1.50‐Å resolution reveals it as a member of the aspartate aminotransferase fold type I (AAT‐I). Structural comparisons of AtmS13 with homologous SATs that act upon similar substrates implicate potential active site residues that contribute to distinctions in sugar C5 (hexose vs. pentose) and/or sugar C2 (deoxy vs. hydroxyl) substrate specificity. Proteins 2015; 83:1547–1554. © 2015 Wiley Periodicals, Inc.  相似文献   

10.
11.
Recent studies have demonstrated that the O‐antigens of some pathogenic bacteria such as Brucella abortus, Francisella tularensis, and Campylobacter jejuni contain quite unusual N‐formylated sugars (3‐formamido‐3,6‐dideoxy‐d ‐glucose or 4‐formamido‐4,6‐dideoxy‐d ‐glucose). Typically, four enzymes are required for the formation of such sugars: a thymidylyltransferase, a 4,6‐dehydratase, a pyridoxal 5'‐phosphate or PLP‐dependent aminotransferase, and an N‐formyltransferase. To date, there have been no published reports of N‐formylated sugars associated with Mycobacterium tuberculosis. A recent investigation from our laboratories, however, has demonstrated that one gene product from M. tuberculosis, Rv3404c, functions as a sugar N‐formyltransferase. Given that M. tuberculosis produces l ‐rhamnose, both a thymidylyltransferase (Rv0334) and a 4,6‐dehydratase (Rv3464) required for its formation have been identified. Thus, there is one remaining enzyme needed for the production of an N‐formylated sugar in M. tuberculosis, namely a PLP‐dependent aminotransferase. Here we demonstrate that the M. tuberculosis rv3402c gene encodes such an enzyme. Our data prove that M. tuberculosis contains all of the enzymatic activities required for the formation of dTDP‐4‐formamido‐4,6‐dideoxy‐d ‐glucose. Indeed, the rv3402c gene product likely contributes to virulence or persistence during infection, though its temporal expression and location remain to be determined.  相似文献   

12.
Campylobacter jejuni is a Gram‐negative bacterium that represents a leading cause of human gastroenteritis worldwide. Of particular concern is the link between C. jejuni infections and the subsequent development of Guillain‐Barré syndrome, an acquired autoimmune disorder leading to paralysis. All Gram‐negative bacteria contain complex glycoconjugates anchored to their outer membranes, but in most strains of C. jejuni, this lipoglycan lacks the O‐antigen repeating units. Recent mass spectrometry analyses indicate that the C. jejuni 81116 (Penner serotype HS:6) lipoglycan contains two dideoxyhexosamine residues, and enzymological assay data show that this bacterial strain can synthesize both dTDP‐3‐acetamido‐3,6‐dideoxy‐d ‐glucose and dTDP‐3‐acetamido‐3,6‐dideoxy‐d ‐galactose. The focus of this investigation is on WlaRG from C. jejuni, which plays a key role in the production of these unusual sugars by functioning as a pyridoxal 5′‐phosphate dependent aminotransferase. Here, we describe the first three‐dimensional structures of the enzyme in various complexes determined to resolutions of 1.7 Å or higher. Of particular significance are the external aldimine structures of WlaRG solved in the presence of either dTDP‐3‐amino‐3,6‐dideoxy‐d ‐galactose or dTDP‐3‐amino‐3,6‐dideoxy‐d ‐glucose. These models highlight the manner in which WlaRG can accommodate sugars with differing stereochemistries about their C‐4′ carbon positions. In addition, we present a corrected structure of WbpE, a related sugar aminotransferase from Pseudomonas aeruginosa, solved to 1.3 Å resolution.  相似文献   

13.
We report herein the crystal structure of Escherichia coli RimK at a resolution of 2.85 Å, an enzyme that catalyzes the post‐translational addition of up to 15 C‐terminal glutamate residues to ribosomal protein S6. The structure belongs to the ATP‐grasp superfamily and is organized as a tetramer, consistent with gel filtration analysis. Each subunit consists of three distinct structural domains and the active site is located in the cleft between these domains. The catalytic reaction appears to occur at the junction between the three domains as ATP binds between the B and C domains, and other substrates bind nearby.Proteins 2013; 81:1847–1854. © 2013 Wiley Periodicals, Inc.  相似文献   

14.
Tuberculosis, caused by the bacterium Mycobacterium tuberculosis, continues to be a major threat to populations worldwide. Whereas the disease is treatable, the drug regimen is arduous at best with the use of four antimicrobials over a six‐month period. There is clearly a pressing need for the development of new therapeutics. One potential target for structure‐based drug design is the enzyme RmlA, a glucose‐1‐phosphate thymidylyltransferase. This enzyme catalyzes the first step in the biosynthesis of l ‐rhamnose, which is a deoxysugar critical for the integrity of the bacterium's cell wall. Here, we report the X‐ray structures of M. tuberculosis RmlA in complex with either dTTP or dTDP‐glucose to 1.6 Å and 1.85 Å resolution, respectively. In the RmlA/dTTP complex, two magnesium ions were observed binding to the nucleotide, both ligated in octahedral coordination spheres. In the RmlA/dTDP‐glucose complex, only a single magnesium ion was observed. Importantly, for RmlA‐type enzymes with known three‐dimensional structures, not one model shows the position of the magnesium ion bound to the nucleotide‐linked sugar. As such, this investigation represents the first direct observation of the manner in which a magnesium ion is coordinated to the RmlA product and thus has important ramifications for structure‐based drug design. In the past, molecular modeling procedures have been employed to derive a three‐dimensional model of the M. tuberculosis RmlA for drug design. The X‐ray structures presented herein provide a superior molecular scaffold for such endeavors in the treatment of one of the world's deadliest diseases.  相似文献   

15.
Antibiotic resistance, along with the resolution of antibiotic-ribosomal subunit complexes at the atomic level, has provided new insights into modifications of clinically relevant antimicrobials that target the ribosome. Modifications to the aminoglycoside or negamycin scaffolds have been reported in the past, but few derivatives appear to be greatly improved compared to their parent compound. Computational and/or traditional screening efforts have yielded novel compounds that bind to the decoding site of the small (30S) ribosomal subunit; naphthyridones appear to bind only in the presence of poly(U) and tRNA(Phe), whereas quinolines bind in a similar manner to aminoglycosides. Streptogramin B analogs were designed that have an amide replacement of the labile ester bond. The resultant molecules were not substrates for the inactivating lyase, but were no longer inhibitors of translation. The synthesis of 16-membered macrolides that are modified at the C6 position with peptidyl moieties as well as conjugates of chloramphenicol to either nucleotide groups or pyrene have been described, but no antibacterial activity has been reported. X-ray crystal structures are now available that can be used to improve on natural or synthetic antibiotics that bind to either the 30S or the 50S ribosomal subunit.  相似文献   

16.
Aminoglycoside has been known as a clinically important antibiotic for a long time, but genetic information for the biosynthesis of aminoglycoside is still insufficient. In this study, we tried to clone aminoglycoside-biosynthetic genes from soil DNA for accumulation of genetic information. We chose the genes encoding L-glutamine:(2-deoxy-)scyllo-inosose aminotransferase as the target, because it is specific for all types of aminoglycoside biosynthesis. By degenerate PCR, we obtained 33 individual clones that were homologous with aminotransferase genes in aminoglycoside biosynthesis. Phylogenetic analysis and alignment of these genes showed that horizontal gene transfer has occurred in the soil. Among these, several quite interesting genes were obtained. Some genes probably originated from non-actinomycetes, and some were far from the known homologs. These genes can be useful markers for the isolation of entire gene clusters and originating organisms.  相似文献   

17.
The four‐subunit protease complex γ‐secretase cleaves many single‐pass transmembrane (TM) substrates, including Notch and β‐amyloid precursor protein to generate amyloid‐β (Aβ), central to Alzheimer's disease. Two of the subunits anterior pharynx‐defective 1 (APH‐1) and presenilin (PS) exist in two homologous forms APH1‐A and APH1‐B, and PS1 and PS2. The consequences of these variations are poorly understood and could affect Aβ production and γ‐secretase medicine. Here, we developed the first complete structural model of the APH‐1B subunit using the published cryo‐electron microscopy (cryo‐EM) structures of APH1‐A (Protein Data Bank: 5FN2, 5A63, and 6IYC). We then performed all‐atom molecular dynamics simulations at 303 K in a realistic bilayer system to understand both APH‐1B alone and in γ‐secretase without and with substrate C83‐bound. We show that APH‐1B adopts a 7TM topology with a water channel topology similar to APH‐1A. We demonstrate direct transport of water through this channel, mainly via Glu84, Arg87, His170, and His196. The apo and holo states closely resemble the experimental cryo‐EM structures with APH‐1A, however with subtle differences: The substrate‐bound APH‐1B γ‐secretase was quite stable, but some TM helices of PS1 and APH‐1B rearranged in the membrane consistent with the disorder seen in the cryo‐EM data. This produces different accessibility of water molecules for the catalytic aspartates of PS1, critical for Aβ production. In particular, we find that the typical distance between the catalytic aspartates of PS1 and the C83 cleavage sites are shorter in APH‐1B, that is, it represents a more closed state, due to interactions with the C‐terminal fragment of PS1. Our structural‐dynamic model of APH‐1B alone and in γ‐secretase suggests generally similar topology but some notable differences in water accessibility which may be relevant to the protein's existence in two forms and their specific function and location.  相似文献   

18.
Kinetic, thermodynamic, and structural properties of the aminoglycoside N3‐acetyltransferase‐VIa (AAC‐VIa) are determined. Among the aminoglycoside N3‐acetyltransferases, AAC‐VIa has one of the most limited substrate profiles. Kinetic studies showed that only five aminoglycosides are substrates for this enzyme with a range of fourfold difference in kcat values. Larger differences in KM (~40‐fold) resulted in ~30‐fold variation in kcat/KM. Binding of aminoglycosides to AAC‐VIa was enthalpically favored and entropically disfavored with a net result of favorable Gibbs energy (ΔG < 0). A net deprotonation of the enzyme, ligand, or both accompanied the formation of binary and ternary complexes. This is opposite of what was observed with several other aminoglycoside N3‐acetyltransferases, where ligand binding causes more protonation. The change in heat capacity (ΔCp) was different in H2O and D2O for the binary enzyme–sisomicin complex but remained the same in both solvents for the ternary enzyme–CoASH–sisomicin complex. Unlike, most other aminoglycoside‐modifying enzymes, the values of ΔCp were within the expected range of protein‐carbohydrate interactions. Solution behavior of AAC‐VIa was also different from the more promiscuous aminoglycoside N3‐acetyltransferases and showed a monomer‐dimer equilibrium as detected by analytical ultracentrifugation (AUC). Binding of ligands shifted the enzyme to monomeric state. Data also showed that polar interactions were the most dominant factor in dimer formation. Overall, thermodynamics of ligand‐protein interactions and differences in protein behavior in solution provide few clues on the limited substrate profile of this enzyme despite its >55% sequence similarity to the highly promiscuous aminoglycoside N3‐acetyltransferase. Proteins 2017; 85:1258–1265. © 2017 Wiley Periodicals, Inc.  相似文献   

19.
Kijanimicin is an antitumor antibiotic isolated from Actinomadura kijaniata. It is composed of three distinct moieties: a pentacyclic core, a monosaccharide referred to as d ‐kijanose, and a tetrasaccharide chain composed of l ‐digitoxose units. d ‐Kijanose is a highly unusual nitro‐containing tetradeoxysugar, which requires at least ten enzymes for its production. Here we describe a structural analysis of one of these enzymes, namely KijD1, which functions as a C‐3′‐methyltransferase using S‐adenosylmethionine as its cofactor. For this investigation, two ternary complexes of KijD1, determined in the presence of S‐adenosylhomocysteine (SAH) and dTDP or SAH and dTDP‐3‐amino‐2,3,6‐trideoxy‐4‐keto‐3‐methyl‐d ‐glucose, were solved to 1.7 or 1.6 Å resolution, respectively. Unexpectedly, these structures, as well as additional biochemical analyses, demonstrated that the quaternary structure of KijD1 is a dimer. Indeed, this is in sharp contrast to that previously observed for the sugar C‐3′‐methyltransferase isolated from Micromonospora chalcea. By the judicious use of site‐directed mutagenesis, it was possible to convert the dimeric form of KijD1 into a monomeric version. The quaternary structure of KijD1 could not have been deduced based solely on bioinformatics approaches, and thus this investigation highlights the continuing need for experimental validation.  相似文献   

20.
Maleylacetate reductase plays a crucial role in catabolism of resorcinol by catalyzing the NAD(P)H‐dependent reduction of maleylacetate, at a carbon–carbon double bond, to 3‐oxoadipate. The crystal structure of maleylacetate reductase from Rhizobium sp. strain MTP‐10005, GraC, has been elucidated by the X‐ray diffraction method at 1.5 Å resolution. GraC is a homodimer, and each subunit consists of two domains: an N‐terminal NADH‐binding domain adopting an α/β structure and a C‐terminal functional domain adopting an α‐helical structure. Such structural features show similarity to those of the two existing families of enzymes in dehydroquinate synthase‐like superfamily. However, GraC is distinct in dimer formation and activity expression mechanism from the families of enzymes. Two subunits in GraC have different structures from each other in the present crystal. One subunit has several ligands mimicking NADH and the substrate in the cleft and adopts a closed domain arrangement. In contrast, the other subunit does not contain any ligand causing structural changes and adopts an open domain arrangement. The structure of GraC reveals those of maleylacetate reductase both in the coenzyme, substrate‐binding state and in the ligand‐free state. The comparison of both subunit structures reveals a conformational change of the Tyr326 loop for interaction with His243 on ligand binding. Structures of related enzymes suggest that His243 is likely a catalytic residue of GraC. Mutational analyses of His243 and Tyr326 support the catalytic roles proposed from structural information. The crystal structure of GraC characterizes the maleylacetate reductase family as a third family in the dehydroquinate synthase‐like superfamily. Proteins 2016; 84:1029–1042. © 2016 Wiley Periodicals, Inc.  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