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.
Cholix toxin from Vibrio cholerae is the third member of the diphtheria toxin (DT) group of mono-ADP-ribosyltransferase (mART) bacterial toxins. It shares structural and functional properties with Pseudomonas aeruginosa exotoxin A and Corynebacterium diphtheriae DT. Cholix toxin is an important model for the development of antivirulence approaches and therapeutics against these toxins from pathogenic bacteria. Herein, we have used the high-resolution X-ray structure of full-length cholix complexed with NAD+ to describe the properties of the NAD+-binding pocket at the residue level, including the role of crystallographic water molecules in the NAD+ substrate interaction. The full-length apo cholix structure is used to describe the putative NAD+-binding site(s) and to correlate biochemical with crystallographic data to study the stoichiometry and orientation of bound NAD+ molecules. We quantitatively describe the NAD+ substrate interactions on a residue basis for the main 22 pocket residues in cholixf, a glycerol and 5 contact water molecules as part of the recognition surface by the substrate according to the conditions of crystallization. In addition, the dynamic properties of an in silico version of the catalytic domain were investigated in order to understand the lack of electronic density for one of the main flexible loops (R-loop) in the pocket of X-ray complexes. Implications for a rational drug design approach for mART toxins are derived.  相似文献   

2.
Certain Vibrio cholerae strains produce cholix, a potent protein toxin that has diphthamide-specific ADP-ribosyltransferase activity against eukaryotic elongation factor 2. Here we present a 1.8 Å crystal structure of cholix in complex with its natural substrate, nicotinamide adenine dinucleotide (NAD+). We also substituted hallmark catalytic residues by site-directed mutagenesis and analyzed both NAD+ binding and ADP-ribosyltransferase activity using a fluorescence-based assay. These data are the basis for a new kinetic model of cholix toxin activity. Further, the new structural data serve as a reference for continuing inhibitor development for this toxin class.  相似文献   

3.
Type II toxin‐antitoxin (TA) modules, which are important cellular regulators in prokaryotes, usually encode two proteins, a toxin that inhibits cell growth and a nontoxic and labile inhibitor (antitoxin) that binds to and neutralizes the toxin. Here, we demonstrate that the res‐xre locus from Photorhabdus luminescens and other bacterial species function as bona fide TA modules in Escherichia coli. The 2.2 Å crystal structure of the intact Pseudomonas putida RES‐Xre TA complex reveals an unusual 2:4 stoichiometry in which a central RES toxin dimer binds two Xre antitoxin dimers. The antitoxin dimers each expose two helix‐turn‐helix DNA‐binding domains of the Cro repressor type, suggesting the TA complex is capable of binding the upstream promoter sequence on DNA. The toxin core domain shows structural similarity to ADP‐ribosylating enzymes such as diphtheria toxin but has an atypical NAD+‐binding pocket suggesting an alternative function. We show that activation of the toxin in vivo causes a depletion of intracellular NAD+ levels eventually leading to inhibition of cell growth in E. coli and inhibition of global macromolecular biosynthesis. Both structure and activity are unprecedented among bacterial TA systems, suggesting the functional scope of bacterial TA toxins is much wider than previously appreciated.  相似文献   

4.
Aldehyde dehydrogenases are found in all organisms and play an important role in the metabolic conversion and detoxification of endogenous and exogenous aldehydes. Genomes of many organisms including Escherichia coli and Salmonella typhimurium encode two succinate semialdehyde dehydrogenases with low sequence similarity and different cofactor preference (YneI and GabD). Here, we present the crystal structure and biochemical characterization of the NAD(P)+‐dependent succinate semialdehyde dehydrogenase YneI from S. typhimurium. This enzyme shows high activity and affinity toward succinate semialdehyde and exhibits substrate inhibition at concentrations of SSA higher than 0.1 mM. YneI can use both NAD+ and NADP+ as cofactors, although affinity to NAD+ is 10 times higher. High resolution crystal structures of YneI were solved in a free state (1.85 Å) and in complex with NAD+ (1.90 Å) revealing a two domain protein with the active site located in the interdomain interface. The NAD+ molecule is bound in the long channel with its nicotinamide ring positioned close to the side chain of the catalytic Cys268. Site‐directed mutagenesis demonstrated that this residue, as well as the conserved Trp136, Glu365, and Asp426 are important for activity of YneI, and that the conserved Lys160 contributes to the enzyme preference to NAD+. Our work has provided further insight into the molecular mechanisms of substrate selectivity and activity of succinate semialdehyde dehydrogenases. © 2012 Wiley Periodicals, Inc.  相似文献   

5.
The crystal structure of the NADH:quinone oxidoreductase PA1024 has been solved in complex with NAD+ to 2.2 Å resolution. The nicotinamide C4 is 3.6 Å from the FMN N5 atom, with a suitable orientation for facile hydride transfer. NAD+ binds in a folded conformation at the interface of the TIM‐barrel domain and the extended domain of the enzyme. Comparison of the enzyme‐NAD+ structure with that of the ligand‐free enzyme revealed a different conformation of a short loop (75–86) that is part of the NAD+‐binding pocket. P78, P82, and P84 provide internal rigidity to the loop, whereas Q80 serves as an active site latch that secures the NAD+ within the binding pocket. An interrupted helix consisting of two α‐helices connected by a small three‐residue loop binds the pyrophosphate moiety of NAD+. The adenine moiety of NAD+ appears to π–π stack with Y261. Steric constraints between the adenosine ribose of NAD+, P78, and Q80, control the strict specificity of the enzyme for NADH. Charged residues do not play a role in the specificity of PA1024 for the NADH substrate.  相似文献   

6.
Domain‐swopped chimeras of the glutamate dehydrogenases from Clostridium symbiosum (CsGDH) (NAD+‐specific) and Escherichia coli (EcGDH) (NADP+‐specific) have been produced, with the aim of testing the localization of determinants of coenzyme specificity. An active chimera consisting of the substrate‐binding domain (Domain I) of CsGDH and the coenzyme‐binding domain (Domain II) of EcGDH has been purified to homogeneity, and a thorough kinetic analysis has been carried out. Results indicate that selectivity for the phosphorylated coenzyme does indeed reside solely in Domain II; the chimera utilizes NAD+ at 0.8% of the rate observed with NADP+, similar to the 0.5% ratio for EcGDH. Positive cooperativity toward L ‐glutamate, characteristic of CsGDH, has been retained with Domain I. An unforeseen feature of this chimera, however, is that, although glutamate cooperativity occurs only at higher pH values in the parent CsGDH, the chimeric protein shows it over the full pH range explored. Also surprising is that the chimera is capable of catalysing severalfold higher reaction rates (Vmax) in both directions than either of the parent enzymes from which it is constructed. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

7.
8.
For the human pathogen Clostridioides (also known as Clostridium) difficile, the ability to adapt to nutrient availability is critical for its proliferation and production of toxins during infection. Synthesis of the toxins is regulated by the availability of certain carbon sources, fermentation products and amino acids (e.g. proline, cysteine, isoleucine, leucine and valine). The effect of proline is attributable at least in part to its role as an inducer and substrate of D‐proline reductase (PR), a Stickland reaction that regenerates NAD+ from NADH. Many Clostridium spp. use Stickland metabolism (co‐fermentation of pairs of amino acids) to generate ATP and NAD+. Synthesis of PR is activated by PrdR, a proline‐responsive regulatory protein. Here we report that PrdR, in the presence of proline, represses other NAD+‐generating pathways, such as the glycine reductase and succinate‐acetyl CoA utilization pathways leading to butyrate production, but does so indirectly by affecting the activity of Rex, a global redox‐sensing regulator that responds to the NAD+/NADH ratio. Our results indicate that PR activity is the favored mechanism for NAD+ regeneration and that both Rex and PrdR influence toxin production. Using the hamster model of C. difficile infection, we revealed the importance of PrdR‐regulated Stickland metabolism in the virulence of C. difficile.  相似文献   

9.
The soluble and calmodulin (CaM)‐dependent NAD+ kinase from Lycopersicon pimpinellifolium was previously shown to be largely inactivated in isolated cells exposed to a short‐term NaCl stress (Delumeau, Morère‐Le Paven, Montrichard, Laval‐Martin (2000) Plant Cell & Environment 23, 329–336). Nevertheless, the activity could be restored by adding a high dithiothreitol concentration to the protein extract, suggesting that the salt stress triggers an oxidation of the enzyme which leads to its inactivation. It was then interesting to investigate the effect of thiol‐modifying reagents and disulphide reductants on the activity of L. pimpinellifolium NAD+ kinase. A three‐step purification procedure was then established and allowed isolation of the enzyme which exists under two forms: a monomer and a dimer of a 56 kDa subunit, characterized, respectively, by pIs of 6·8 and 7·1. Isolated NAD+ kinase had a high affinity for CaM, half saturation being obtained for 7 ng mL?1 bovine CaM. The activity of NAD+ kinase was strongly inhibited by thiol‐modifying reagents and oxidized glutathione. NAD+ kinase was also found to be air‐inactivated, the residual activity being stimulated by disulphide reductants. The most efficient of them is reduced thioredoxin from Escherichia coli which induced a five‐fold increase in activity and restored 80% of the initial activity. These results which can be related to those previously observed in vivo suggest that the activity of the L. pimpinellifolium NAD+ kinase, besides its dependence on CaM, is also dependent on the reduction state of the protein which could be regulated by the thioredoxin h/NADP‐thioredoxin reductase system.  相似文献   

10.
The mono-ADP-ribosyltransferase (mART) toxins are contributing factors to a number of human diseases, including cholera, diphtheria, traveler''s diarrhea, and whooping cough. VahC is a cytotoxic, actin-targeting mART from Aeromonas hydrophila PPD134/91. This bacterium is implicated primarily in diseases among freshwater fish species but also contributes to gastrointestinal and extraintestinal infections in humans. VahC was shown to ADP-ribosylate Arg-177 of actin, and the kinetic parameters were Km(NAD+) = 6 μm, Km(actin) = 24 μm, and kcat = 22 s−1. VahC activity caused depolymerization of actin filaments, which induced caspase-mediated apoptosis in HeLa Tet-Off cells. Alanine-scanning mutagenesis of predicted catalytic residues showed the predicted loss of in vitro mART activity and cytotoxicity. Bioinformatic and kinetic analysis also identified three residues in the active site loop that were critical for the catalytic mechanism. A 1.9 Å crystal structure supported the proposed roles of these residues and their conserved nature among toxin homologues. Several small molecules were characterized as inhibitors of in vitro VahC mART activity and suramin was the best inhibitor (IC50 = 20 μm). Inhibitor activity was also characterized against two other actin-targeting mART toxins. Notably, these inhibitors represent the first report of broad spectrum inhibition of actin-targeting mART toxins.  相似文献   

11.
12.
13.
Interventions that slow aging and prevent chronic disease may come from an understanding of how dietary restriction (DR) increases lifespan. Mechanisms proposed to mediate DR longevity include reduced mTOR signaling, activation of the NAD+‐dependent deacylases known as sirtuins, and increases in NAD+ that derive from higher levels of respiration. Here, we explored these hypotheses in Caenorhabditis elegans using a new liquid feeding protocol. DR lifespan extension depended upon a group of regulators that are involved in stress responses and mTOR signaling, and have been implicated in DR by some other regimens [DAF‐16 (FOXO), SKN‐1 (Nrf1/2/3), PHA‐4 (FOXA), AAK‐2 (AMPK)]. Complete DR lifespan extension required the sirtuin SIR‐2.1 (SIRT1), the involvement of which in DR has been debated. The nicotinamidase PNC‐1, a key NAD+ salvage pathway component, was largely required for DR to increase lifespan but not two healthspan indicators: movement and stress resistance. Independently of pnc‐1, DR increased the proportion of respiration that is coupled to ATP production but, surprisingly, reduced overall oxygen consumption. We conclude that stress response and NAD+‐dependent mechanisms are each critical for DR lifespan extension, although some healthspan benefits do not require NAD+ salvage. Under DR conditions, NAD+‐dependent processes may be supported by a DR‐induced shift toward oxidative metabolism rather than an increase in total respiration.  相似文献   

14.
To elucidate the molecular mechanisms underlying non‐alcoholic fatty liver disease (NAFLD), we recruited 86 subjects with varying degrees of hepatic steatosis (HS). We obtained experimental data on lipoprotein fluxes and used these individual measurements as personalized constraints of a hepatocyte genome‐scale metabolic model to investigate metabolic differences in liver, taking into account its interactions with other tissues. Our systems level analysis predicted an altered demand for NAD+ and glutathione (GSH) in subjects with high HS. Our analysis and metabolomic measurements showed that plasma levels of glycine, serine, and associated metabolites are negatively correlated with HS, suggesting that these GSH metabolism precursors might be limiting. Quantification of the hepatic expression levels of the associated enzymes further pointed to altered de novo GSH synthesis. To assess the effect of GSH and NAD+ repletion on the development of NAFLD, we added precursors for GSH and NAD+ biosynthesis to the Western diet and demonstrated that supplementation prevents HS in mice. In a proof‐of‐concept human study, we found improved liver function and decreased HS after supplementation with serine (a precursor to glycine) and hereby propose a strategy for NAFLD treatment.  相似文献   

15.
Age‐related hearing loss (ARHL) is a major neurodegenerative disorder and the leading cause of communication deficit in the elderly population, which remains largely untreated. The development of ARHL is a multifactorial event that includes both intrinsic and extrinsic factors. Recent studies suggest that NAD+/NADH ratio may play a critical role in cellular senescence by regulating sirtuins, PARP‐1, and PGC‐1α. Nonetheless, the beneficial effect of direct modulation of cellular NAD+ levels on aging and age‐related diseases has not been studied, and the underlying mechanisms remain obscure. Herein, we investigated the effect of β‐lapachone (β‐lap), a known plant‐derived metabolite that modulates cellular NAD+ by conversion of NADH to NAD+ via the enzymatic action of NADH: quinone oxidoreductase 1 (NQO1) on ARHL in C57BL/6 mice. We elucidated that the reduction of cellular NAD+ during the aging process was an important contributor for ARHL; it facilitated oxidative stress and pro‐inflammatory responses in the cochlear tissue through regulating sirtuins that alter various signaling pathways, such as NF‐κB, p53, and IDH2. However, augmentation of NAD+ by β‐lap effectively prevented ARHL and accompanying deleterious effects through reducing inflammation and oxidative stress, sustaining mitochondrial function, and promoting mitochondrial biogenesis in rodents. These results suggest that direct regulation of cellular NAD+ levels by pharmacological agents may be a tangible therapeutic option for treating various age‐related diseases, including ARHL.  相似文献   

16.
The NAD+‐dependent lactate dehydrogenase from Bacillus subtilis (BsLDH) catalyzes the enantioselective reduction of pyruvate to lactate. BsLDH is highly specific to NAD+ and exhibits only a low activity with NADP+ as cofactor. Based on the high activity and good stability of LDHs, these enzymes have been frequently used for the regeneration of NAD+. While an application in the regeneration of NADP+ is not sufficient due to the cofactor preference of the BsLDH. In addition, NADP+‐dependent LDHs have not yet been found in nature. Therefore, a structure‐based approach was performed to predict amino acids involved in the cofactor specificity. Methods of site‐saturation mutagenesis were applied to vary these amino acids, with the aim to alter the cofactor specificity of the BsLDH. Five constructed libraries were screened for improved NADP+ acceptance. The mutant V39R was identified to have increased activity with NADP+ relative to the wild type. V39R was purified and biochemically characterized. V39R showed excellent kinetic properties with NADP(H) and NAD(H), for instance the maximal specific activity with NADPH was enhanced 100‐fold to 90.8 U/mg. Furthermore, a 249‐fold increased catalytic efficiency was observed. Surprisingly, the activity with NADH was also significantly improved. Overall, we were able to successfully apply V39R in the regeneration of NADP+ in an enzyme‐coupled approach combined with the NADP+‐dependent alcohol dehydrogenase from Lactobacillus kefir. We demonstrate for the first time an application of an LDH in the regeneration of NADP+.  相似文献   

17.
We identified Certhrax, the first anthrax-like mART toxin from the pathogenic G9241 strain of Bacillus cereus. Certhrax shares 31% sequence identity with anthrax lethal factor from Bacillus anthracis; however, we have shown that the toxicity of Certhrax resides in the mART domain, whereas anthrax uses a metalloprotease mechanism. Like anthrax lethal factor, Certhrax was found to require protective antigen for host cell entry. This two-domain enzyme was shown to be 60-fold more toxic to mammalian cells than anthrax lethal factor. Certhrax localizes to distinct regions within mouse RAW264.7 cells by 10 min postinfection and is extranuclear in its cellular location. Substitution of catalytic residues shows that the mART function is responsible for the toxicity, and it binds NAD+ with high affinity (KD = 52.3 ± 12.2 μm). We report the 2.2 Å Certhrax structure, highlighting its structural similarities and differences with anthrax lethal factor. We also determined the crystal structures of two good inhibitors (P6 (KD = 1.7 ± 0.2 μm, Ki = 1.8 ± 0.4 μm) and PJ34 (KD = 5.8 ± 2.6 μm, Ki = 9.6 ± 0.3 μm)) in complex with Certhrax. As with other toxins in this family, the phosphate-nicotinamide loop moves toward the NAD+ binding site with bound inhibitor. These results indicate that Certhrax may be important in the pathogenesis of B. cereus.  相似文献   

18.
Betaine aldehyde dehydrogenase (BADH EC 1.2.1.8) catalyzes the irreversible oxidation of betaine aldehyde to glycine betaine using NAD+ as a coenzyme. Porcine kidney BADH (pkBADH) follows a bi‐bi ordered mechanism in which NAD+ binds to the enzyme before the aldehyde. Previous studies showed that NAD+ induces complex and unusual conformational changes on pkBADH and that potassium is required to maintain its quaternary structure. The aim of this work was to analyze the structural changes in pkBADH caused by NAD+ binding and the role played by potassium in those changes. The pkBADH cDNA was cloned and overexpressed in Escherichia coli, and the protein was purified by affinity chromatography using a chitin matrix. The pkBADH/NAD+ interaction was analyzed by circular dichroism (CD) and by isothermal titration calorimetry (ITC) by titrating the enzyme with NAD+. The cDNA has an open reading frame of 1485 bp and encodes a protein of 494 amino acids, with a predicted molecular mass of 53.9 kDa. CD data showed that the binding of NAD+ to the enzyme caused changes in its secondary structure, whereas the presence of K+ helps maintain its α‐helix content. K+ increased the thermal stability of the pkBADH‐NAD+ complex by 5.3°C. ITC data showed that NAD+ binding occurs with different association constants for each active site between 37.5 and 8.6 μM. All the results support previous data in which the enzyme incubation with NAD+ provoked changes in reactivity, which is an indication of slow conformational rearrangements of the active site.  相似文献   

19.
Cyclic ADP‐ribose (cADPR) mobilizes intracellular Ca2+ stores and activates Ca2+ influx to regulate a wide range of physiological processes. It is one of the products produced from the catalysis of NAD+ by the multifunctional CD38/ADP‐ribosyl cyclase superfamily. After elimination of the nicotinamide ring by the enzyme, the reaction intermediate of NAD+ can either be hydrolyzed to form linear ADPR or cyclized to form cADPR. We have previously shown that human CD38 exhibits a higher preference towards the hydrolysis of NAD+ to form linear ADPR while Aplysia ADP‐ribosyl cyclase prefers cyclizing NAD+ to form cADPR. In this study, we characterized the enzymatic properties of porcine CD38 and revealed that it has a prominent secondary NAD+ cyclase activity producing cADPR. We also determined the X‐ray crystallographic structures of porcine CD38 and were able to observe conformational flexibility at the base of the active site of the enzyme which allow the NAD+ reaction intermediate to adopt conformations resulting in both hydrolysis and cyclization forming linear ADPR and cADPR respectively.  相似文献   

20.
A 50-ns molecular dynamics simulation has been used to study the homotetramer of the enzyme glycosomal glyceraldehyde 3-phosphate dehydrogenase (gGAPDH) complexes, from Trypanosoma cruzi, with nicotinamide adenine dinucleotide (NAD+) cofactors in aqueous solution. The root mean square deviation indicates that the overall structure of the homotetramer does not undergo significant change. The largest structural change observed was in the NAD+ binding domain of subunit (chain) D; as a consequence, the NAD+ cofactor was dislocated from its initial position. However, the other subunits were not affected, suggesting that the gGAPDH enzyme exhibits non-cooperative behaviour. Our simulation estimates that the NAD+ binding domain rotates about 4.8° relative to the catalytic domain in the apo–holo form transition. The hydrogen bond analysis reveals that the residues R12, I13, D38 and M39 are essential for gGAPDH–NAD+ interaction. Furthermore, two promising cavities to be explored in drug design were found: one formed by residues I13, R12, T197, T199, E336 and Y339, and the other by residues C166, H194, R249, I13, R12, T197, T199, E336 and Y339. The results presented in this paper offer new insight into the search for inhibitors of the gGAPDH enzyme of T. cruzi protozoan.  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