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.
Nitration of tyrosine residues has been shown to be an important oxidative modification in proteins and has been suggested to play a role in several diseases such as atherosclerosis, asthma, lung and neurodegenerative diseases. Detection of nitrated proteins has been mainly based on the use of nitrotyrosine‐specific antibodies. In contrast, only a small number of nitration sites in proteins have been unequivocally identified by MS. We have used a monoclonal 3‐NT‐specific antibody, and have synthesized a series of tyrosine‐nitrated peptides of prostacyclin synthase (PCS) in which a single specific nitration site at Tyr‐430 had been previously identified upon reaction with peroxynitrite 17 . The determination of antibody‐binding affinity and specificity of PCS peptides nitrated at different tyrosine residues (Tyr‐430, Tyr‐421, Tyr‐83) and sequence mutations around the nitration sites provided the identification of an epitope motif containing positively charged amino acids (Lys and/or Arg) N‐terminal to the nitration site. The highest affinity to the anti‐3NT‐antibody was found for the PCS peptide comprising the Tyr‐430 nitration site with a KD of 60 nM determined for the peptide, PCS(424‐436‐Tyr‐430NO2); in contrast, PCS peptides nitrated at Tyr‐421 and Tyr‐83 had substantially lower affinity. ELISA, SAW bioaffinity, proteolytic digestion of antibody‐bound peptides and affinity‐MS analysis revealed highest affinity to the antibody for tyrosine‐nitrated peptides that contained positively charged amino acids in the N‐terminal sequence to the nitration site. Remarkably, similar N‐terminal sequences of tyrosine‐nitration sites have been recently identified in nitrated physiological proteins, such as eosinophil peroxidase and eosinophil‐cationic protein. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
3.
4.
Cholix toxin from Vibrio cholerae is a novel mono‐ADP‐ribosyltransferase (mART) toxin that shares structural and functional properties with Pseudomonas aeruginosa exotoxin A and Corynebacterium diphtheriae diphtheria toxin. Herein, we have used the high‐resolution X‐ray structure of full‐length cholix toxin in the apo form, NAD+ bound, and 10 structures of the cholix catalytic domain (C‐domain) complexed with several strong inhibitors of toxin enzyme activity (NAP, PJ34, and the P‐series) to study the binding mode of the ligands. A pharmacophore model based on the active pose of NAD+ was compared with the active conformation of the inhibitors, which revealed a cationic feature in the side chain of the inhibitors that may determine the active pose. Moreover, a conformational search was conducted for the missing coordinates of one of the main active‐site loops (R‐loop). The resulting structural models were used to evaluate the interaction energies and for 3D‐QSAR modeling. Implications for a rational drug design approach for mART toxins were derived. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

5.
AIM:To characterize phosphorylation of human glyceraldehyde 3-phosphate dehydrogenase(GAPDH),and mobility of GAPDH in cancer cells treated with chemotherapeutic agents. METHODS:We used proteomics analysis to detect and characterize phosphorylation sites within human GAPDH. Site-specific mutagenesis and alanine scanning was then performed to evaluate functional significance of phosphorylation sites in the GAPDH polypeptide chain. Enzymatic properties of mutated GAPDH variants were assessed using kinetic studies. Intranuclear dynamics parameters(diffusion coefficient and the immobile fraction) were estimated using fluorescence recovery after photobleaching(FRAP) experiments and confocal microscopy. Molecular modeling experiments were performed to estimate the effects of mutations on NAD+ cofactor binding.RESULTS:Using MALDI-TOF analysis,we identified novel phosphorylation sites within the NAD+ binding center of GAPDH at Y94,S98,and T99. Using polyclonal antibody specific to phospho-T99-containing peptide within GAPDH,we demonstrated accumulation of phospho-T99-GAPDH inthe nuclear fractions of A549,HCT116,and SW48 cancer cel s after cytotoxic stress. We performed site-mutagenesis,and estimated enzymatic properties,intranuclear distribution,and intranuclear mobility of GAPDH mutated variants. Site-mutagenesis at positions S98 and T99 in the NAD+ binding center reduced enzymatic activity of GAPDH due to decreased affinity to NAD+(Km = 741 ± 257 μmol/L in T99 I vs 57 ± 11.1 μmol/L in wild type GAPDH. Molecular modeling experiments revealed the effect of mutations on NAD+ binding with GAPDH. FRAP(fluorescence recovery after photo bleaching) analysis showed that mutations in NAD+ binding center of GAPDH abrogated its intranuclear interactions. CONCLUSION:Our results suggest an important functional role of phosphorylated amino acids in the NAD+ binding center in GAPDH interactions with its intranuclear partners.  相似文献   

6.
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   

7.
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   

8.
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   

9.
The properties of the active center of phosphorylating D-glyceraldehyde-3-phosphate dehydrogenase (GAPDH) are considered with emphasis on the structure of anion-binding sites and their role in catalysis. The results of studies on the molecular mechanism of the effect of NAD+ on the enzyme conformation are discussed. Experimental evidence is presented supporting the idea that negative cooperativity of NAD+ binding and half-of-the-sites reactivity exhibited by GAPDH are generated by different mechanisms. Data obtained with rabbit muscle and Escherichia coli GAPDH point to preexisting asymmetry in these tetramers. Structural determinants that can control the transition of the tetramer from the symmetric to the asymmetric state were found.  相似文献   

10.
Lower plant species including some green algae, non‐vascular plants (bryophytes) as well as the oldest vascular plants (lycopods) and ferns (monilophytes) possess a unique aldehyde dehydrogenase (ALDH) gene named ALDH21, which is upregulated during dehydration. However, the gene is absent in flowering plants. Here, we show that ALDH21 from the moss Physcomitrella patens codes for a tetrameric NADP+‐dependent succinic semialdehyde dehydrogenase (SSALDH), which converts succinic semialdehyde, an intermediate of the γ‐aminobutyric acid (GABA) shunt pathway, into succinate in the cytosol. NAD+ is a very poor coenzyme for ALDH21 unlike for mitochondrial SSALDHs (ALDH5), which are the closest related ALDH members. Structural comparison between the apoform and the coenzyme complex reveal that NADP+ binding induces a conformational change of the loop carrying Arg‐228, which seals the NADP+ in the coenzyme cavity via its 2′‐phosphate and α‐phosphate groups. The crystal structure with the bound product succinate shows that its carboxylate group establishes salt bridges with both Arg‐121 and Arg‐457, and a hydrogen bond with Tyr‐296. While both arginine residues are pre‐formed for substrate/product binding, Tyr‐296 moves by more than 1 Å. Both R121A and R457A variants are almost inactive, demonstrating a key role of each arginine in catalysis. Our study implies that bryophytes but presumably also some green algae, lycopods and ferns, which carry both ALDH21 and ALDH5 genes, can oxidize SSAL to succinate in both cytosol and mitochondria, indicating a more diverse GABA shunt pathway compared with higher plants carrying only the mitochondrial ALDH5.  相似文献   

11.
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   

12.
Tat is a multifunctional transactivator encoded by human immunodeficiency virus type 1 (HIV‐1). Tat transactivating activity is controlled by nicotinamide adenine nucleotide+ (NAD+)‐dependent deacetylase sirtuin 1 (SIRT1). Nicotinamide phosphoribosyltransferase (Nampt) is a rate‐limiting enzyme in the conversion of nicotinamide into NAD+, which is crucial for SIRT1 activation. Thus, the effect of Nampt on Tat‐regulated SIRT activity was studied in Hela‐CD4‐β‐gal (MAGI) cells. We demonstrated that Tat caused NAD+ depletion and inhibited Nampt mRNA and protein expression in MAGI cells. Resveratrol reversed Tat‐induced NAD+ depletion and inhibition of Nampt mRNA and protein expression. Further investigation revealed that Tat‐induced inhibition of SIRT1 activity was potentiated in Nampt‐knockdown by Nampt siRNA compared to treatment with Tat alone. Nampt siRNA potentiated Tat‐induced HIV‐1 transactivation in MAGI cells. Altogether, these results indicate that Nampt is critical in the regulation of Tat‐induced inhibition of SIRT1 activity and long terminal repeat (LTR) transactivation. Nampt/SIRT1 pathway could be a novel therapeutic tool for the treatment of HIV‐1 infection. J. Cell. Biochem. 110: 1464–1470, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

13.
14.
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   

15.
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   

16.
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   

17.
Despite significant influence of secondary bile acids on human health and disease, limited structural and biochemical information is available for the key gut microbial enzymes catalyzing its synthesis. Herein, we report apo‐ and cofactor bound crystal structures of BaiA2, a short chain dehydrogenase/reductase from Clostridium scindens VPI 12708 that represent the first protein structure of this pathway. The structures elucidated the basis of cofactor specificity and mechanism of proton relay. A conformational restriction involving Glu42 located in the cofactor binding site seems crucial in determining cofactor specificity. Limited flexibility of Glu42 results in imminent steric and electrostatic hindrance with 2′‐phosphate group of NADP(H). Consistent with crystal structures, steady state kinetic characterization performed with both BaiA2 and BaiA1, a close homolog with 92% sequence identity, revealed specificity constant (kcat/KM) of NADP+ at least an order of magnitude lower than NAD+. Substitution of Glu42 with Ala improved specificity toward NADP+ by 10‐fold compared to wild type. The cofactor bound structure uncovered a novel nicotinamide‐hydroxyl ion (NAD+‐OH?) adduct contraposing previously reported adducts. The OH? of the adduct in BaiA2 is distal to C4 atom of nicotinamide and proximal to 2′‐hydroxyl group of the ribose moiety. Moreover, it is located at intermediary distances between terminal functional groups of active site residues Tyr157 (2.7 Å) and Lys161 (4.5 Å). Based on these observations, we propose an involvement of NAD+‐OH? adduct in proton relay instead of hydride transfer as noted for previous adducts. Proteins 2014; 82:216–229. © 2013 Wiley Periodicals, Inc.  相似文献   

18.
Protein tyrosine (Tyr) nitration is a post‐translational modification yielding 3‐nitrotyrosine (NO2–Tyr). Formation of NO2–Tyr is generally considered as a marker of nitro‐oxidative stress and is involved in some human pathophysiological disorders, but has been poorly studied in plants. Leghemoglobin (Lb) is an abundant hemeprotein of legume nodules that plays an essential role as an O2 transporter. Liquid chromatography coupled to tandem mass spectrometry was used for a targeted search and quantification of NO2–Tyr in Lb. For all Lbs examined, Tyr30, located in the distal heme pocket, is the major target of nitration. Lower amounts were found for NO2–Tyr25 and NO2–Tyr133. Nitrated Lb and other as yet unidentified nitrated proteins were also detected in nodules of plants not receiving and were found to decrease during senescence. This demonstrates formation of nitric oxide (˙NO) and by alternative means to nitrate reductase, probably via a ˙NO synthase‐like enzyme, and strongly suggests that nitrated proteins perform biological functions and are not merely metabolic byproducts. In vitro assays with purified Lb revealed that Tyr nitration requires  + H2O2 and that peroxynitrite is not an efficient inducer of nitration, probably because Lb isomerizes it to . Nitrated Lb is formed via oxoferryl Lb, which generates nitrogen dioxide and tyrosyl radicals. This mechanism is distinctly different from that involved in heme nitration. Formation of NO2–Tyr in Lb is a consequence of active metabolism in functional nodules, where Lb may act as a sink of toxic peroxynitrite and may play a protective role in the symbiosis.  相似文献   

19.
We report a combined approach that introduces the use of 4‐aminobenzo‐15‐crown‐5 (4AB15C5) for the detection of ferric(III) ions by colorimetric, ultraviolet (UV)–visible light absorption, fluorescence, and live‐cell imaging techniques along with density functional theory (DFT) calculations. We have found that 4AB15C5 is sensitive and selective for binding ferric(III) ions in aqueous solutions. DFT calculations using the polarizable continuum model have been used to explain the strong binding of the ferric ion by 4AB15C5 in aqueous solutions. The detection limit in the fluorescence quenching measurements was found to be as low as 50 μM for the ferric ion with a determined Stern–Volmer constant of 1.52 × 104 M?1. Fluorescence intensity did not change for other ions tested, Fe2+, Co2+, Mn2+, Mg2+, Zn2+, Ca2+, NH4+, Na+, and K+ ions. Live‐cell fluorescence imaging was also used to check the intracellular variations in ferric ion levels. Our spectroscopic data indicated that 4AB15C5 can bind ferric ions selectively in aqueous solutions.  相似文献   

20.
Mycobacterium tuberculosis catalase‐peroxidase (KatG) is a bifunctional hemoprotein that has been shown to activate isoniazid (INH), a pro‐drug that is integral to frontline antituberculosis treatments. The activated species, presumed to be an isonicotinoyl radical, couples to NAD+/NADH forming an isoniazid‐NADH adduct that ultimately confers anti‐tubercular activity. To better understand the mechanisms of isoniazid activation as well as the origins of KatG‐derived INH‐resistance, we have compared the catalytic properties (including the ability to form the INH‐NADH adduct) of the wild‐type enzyme to 23 KatG mutants which have been associated with isoniazid resistance in clinical M. tuberculosis isolates. Neither catalase nor peroxidase activities, the two inherent enzymatic functions of KatG, were found to correlate with isoniazid resistance. Furthermore, catalase function was lost in mutants which lacked the Met‐Tyr‐Trp crosslink, the biogenic cofactor in KatG which has been previously shown to be integral to this activity. The presence or absence of the crosslink itself, however, was also found to not correlate with INH resistance. The KatG resistance‐conferring mutants were then assayed for their ability to generate the INH‐NADH adduct in the presence of peroxide (t‐BuOOH and H2O2), superoxide, and no exogenous oxidant (air‐only background control). The results demonstrate that residue location plays a critical role in determining INH‐resistance mechanisms associated with INH activation; however, different mutations at the same location can produce vastly different reactivities that are oxidant‐specific. Furthermore, the data can be interpreted to suggest the presence of a second mechanism of INH‐resistance that is not correlated with the formation of the INH‐NADH adduct.  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