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 93 毫秒
1.
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   

2.
CD38 is a type II glycoprotein that catalyzes the formation of cyclic ADP-ribose (cADPR), an intracellular calcium signalling molecule, from nicotinamide adenine dinucleotide (NAD+). Using a modified version of the fluorimetric cycling assay for cADPR which reduces between-subject variability, we report significant decreases in brain and lung cADPR, which although similar to previously published values, showed much less individual variation. The reduced variation within each group suggests that the range of cADPR is narrower than previously thought, and that the regulatory mechanisms controlling these levels are more finely tuned. We also report significant increases in brain, lung, and kidney NAD+ in the Cd38−/− mouse, and provide the first experimental demonstration of the proximate relationship between CD38 and NAD+.  相似文献   

3.
NAD(P)(+)-glycohydrolase (NADase, EC 3.2.2.6) was partially purified from microsomal membranes of human spleen after solubilization with Triton X-100. In addition to NAD+ and NADP+, the enzyme catalyzed the hydrolysis of several NAD+ analogues and the pyridine base exchange reaction with conversion of NAD+ into 3-acetylpyridine adenine dinucleotide. The enzyme also catalyzed the synthesis of cyclic ADP-ribose (cADPR) from NAD+ and the hydrolysis of cADPR to adenosine diphosphoribose (ADPR). Therefore, this enzyme is a new member of multicatalytic NADases recently identified from mammals, involved in the regulation of intracellular cADPR concentration. Human spleen NADase showed a subunit molecular mass of 45 kDa, a pI of 4.9 and a Km value for NAD+ of 26 microM. High activation of ADPR cyclase activity was observed in the presence of Ag+ ions, corresponding to NADase inhibition.  相似文献   

4.
5.
CD38 is a type-II transmembrane glycoprotein occurring in several hematopoietic and mature blood cells as well as in other cell types, including neurons. Although classified as an orphan receptor, CD38 is also a bifunctional ectoenzyme that catalyzes both the conversion of NAD+ to nicotinamide and cyclic ADP-ribose (cADPR), via an ADP-ribosyl cyclase reaction, and also the hydrolysis of cADPR to ADP-ribose (hydrolase). Major unresolved questions concern the correlation between receptor and catalytic properties of CD38, and also the apparent contradiction between ectocellular generation and intracellular Ca2+-mobilizing activity of cADPR. Results are presented that provide some explanations to this topological paradox in two different cell types. In cultured rat cerebellar granule neurons, extracellular cADPR (either generated by CD38 or directly added) elicited an enhanced intracellular Ca2+ response to KCl-induced depolarization, a process that can be qualified as a Ca2+-induced Ca2+ release (CICR) mechanism. On the other hand, in the CD38+ human Namalwa B lymphoid cells, NAD+ (and thiol compounds as well) induced a two-step process of self-aggregation followed by endocytosis of CD38, which resulted in a shift of cADPR metabolism from the cell surface to the cytosol. Both distinctive types of cellular responses to extracellular NAD+ seem to be suitable to elicit changes in the intracellular Ca2+ homeostasis.  相似文献   

6.
During axon degeneration, NAD+ levels are largely controlled by two enzymes: nicotinamide mononucleotide adenylyltransferase 2 (NMNAT2) and sterile alpha and toll interleukin motif containing protein 1 (SARM1). NMNAT2, which catalyzes the formation of NAD+ from NMN and ATP, is actively degraded leading to decreased NAD+ levels. SARM1 activity further decreases the concentration of NAD+ by catalyzing its hydrolysis to form nicotinamide and a mixture of ADPR and cADPR. Notably, SARM1 knockout mice show decreased neurodegeneration in animal models of axon degeneration, highlighting the therapeutic potential of targeting this novel NAD+ hydrolase. This review discusses recent advances in the SARM1 field, including SARM1 structure, regulation, and catalysis as well as the identification of the first SARM1 inhibitors.  相似文献   

7.
SARM1 is the founding member of the TIR-domain family of NAD+ hydrolases and the central executioner of pathological axon degeneration. SARM1-dependent degeneration requires NAD+ hydrolysis. Prior to the discovery that SARM1 is an enzyme, SARM1 was studied as a TIR-domain adaptor protein with non-degenerative signaling roles in innate immunity and invertebrate neurodevelopment, including at the Drosophila neuromuscular junction (NMJ). Here we explore whether the NADase activity of SARM1 also contributes to developmental signaling. We developed transgenic Drosophila lines that express SARM1 variants with normal, deficient, and enhanced NADase activity and tested their function in NMJ development. We find that NMJ overgrowth scales with the amount of NADase activity, suggesting an instructive role for NAD+ hydrolysis in this developmental signaling pathway. While degenerative and developmental SARM1 signaling share a requirement for NAD+ hydrolysis, we demonstrate that these signals use distinct upstream and downstream mechanisms. These results identify SARM1-dependent NAD+ hydrolysis as a heretofore unappreciated component of developmental signaling. SARM1 now joins sirtuins and Parps as enzymes that regulate signal transduction pathways via mechanisms that involve NAD+ cleavage, greatly expanding the potential scope of SARM1 TIR NADase functions.  相似文献   

8.
Streptococcus pyogenes uses the cytolysin streptolysin O (SLO) to translocate an enzyme, the S. pyogenes NAD+ glycohydrolase (SPN), into the host cell cytosol. However, the function of SPN in this compartment is not known. As a complication, many S. pyogenes strains express a SPN variant lacking NAD+ glycohydrolase (NADase) activity. Here, we show that SPN modifies several SLO‐ and NAD+‐dependent host cell responses in patterns that correlate with NADase activity. SLO pore formation results in hyperactivation of the cellular enzyme poly‐ADP‐ribose polymerase‐1 (PARP‐1) and production of polymers of poly‐ADP‐ribose (PAR). However, while SPN NADase activity moderates PARP‐1 activation and blocks accumulation of PAR, these processes continued unabated in the presence of NADase‐inactive SPN. Temporal analyses revealed that while PAR production is initially independent of NADase activity, PAR rapidly disappears in the presence of NADase‐active SPN, host cell ATP is depleted and the pro‐inflammatory mediator high‐mobility group box‐1 (HMGB1) protein is released from the nucleus by a PARP‐1‐dependent mechanism. In contrast, HMGB1 is not released in response to NADase‐inactive SPN and instead the cells release elevated levels of interleukin‐8 and tumour necrosis factor‐α. Thus, SPN and SLO combine to induce cellular responses subsequently influenced by the presence or absence of NADase activity.  相似文献   

9.
CD38 is a type II transmembrane glycoprotein that is expressed by many cell types including lymphocytes. Signaling through CD38 on B lymphocytes can mediate B cell activation, proliferation, and cytokine secretion. Additionally, coligation of CD38 and the B cell Ag receptor can greatly augment B cell Ag receptor responses. Interestingly, the extracellular domain of CD38 catalyzes the conversion of NAD+ into nicotinamide, ADP-ribose (ADPR), and cyclic ADPR (cADPR). cADPR can induce intracellular calcium release in an inositol trisphosphate-independent manner and has been hypothesized to regulate CD38-mediated signaling. We demonstrate that replacement of the cytoplasmic tail and the transmembrane domains of CD38 did not impair CD38 signaling, coreceptor activity, or enzyme activity. In contrast, independent point mutations in the extracellular domain of CD38 dramatically impaired signal transduction. However, no correlation could be found between CD38-mediated signaling and the capacity of CD38 to catalyze an enzyme reaction and produce cADPR, ADPR, and/or nicotinamide. Instead, we propose that CD38 signaling and coreceptor activity in vitro are regulated by conformational changes induced in the extracellular domain upon ligand/substrate binding, rather than on actual turnover or generation of products.  相似文献   

10.
Cyclic ADP-ribose (cADPR) is a universal calcium messenger molecule that regulates many physiological processes. The production and degradation of cADPR are catalyzed by a family of related enzymes, including the ADP-ribosyl cyclase from Aplysia california (ADPRAC) and CD38 from human. Although ADPRC and CD38 share a common evolutionary ancestor, their enzymatic functions toward NAD and cADPR homeostasis have evolved divergently. Thus, ADPRC can only generate cADPR from NAD (cyclase), whereas CD38, in contrast, has multiple activities, i.e. in cADPR production and degradation, as well as NAD hydrolysis (NADase). In this study, we determined a number of ADPRC and CD38 structures bound with various nucleotides. From these complexes, we elucidated the structural features required for the cyclization (cyclase) reaction of ADPRC and the NADase reaction of CD38. Using the structural approach in combination with site-directed mutagenesis, we identified Phe-174 in ADPRC as a critical residue in directing the folding of the substrate during the cyclization reaction. Thus, a point mutation of Phe-174 to glycine can turn ADPRC from a cyclase toward an NADase. The equivalent residue in CD38, Thr-221, is shown to disfavor the cyclizing folding of the substrate, resulting in NADase being the dominant activity. The comprehensive structural comparison of CD38 and APDRC presented in this study thus provides insights into the structural determinants for the functional evolution from a cyclase to a hydrolase.Cyclic ADP-ribose (cADPR)3 is a calcium messenger ubiquitous in mammals as well as in invertebrates and plants and is responsible for regulating many physiological processes ranging from the simple function of calcium channel operation to the complex higher level organization of hormone secretion and autism (reviewed in Lee (1), Schuber and Lund (2), and Malavasi et al. (3)). The enzymatic production of cADPR from the substrate nicotinamide adenine dinucleotide (NAD) requires first the removal of the nicotinamide moiety followed by a cyclization reaction in which both ends of the remaining nucleotide are annealed (Fig. 1A). ADP-ribosyl cyclase (ADPRC) from Aplysia california was the first enzyme discovered to possess this function (cyclase) (4). Based on sequence homology (5), two human antigens, CD38 and CD157, were identified to also have the cyclase activity (68). However, different from ADPRC, which produces only cADPR from NAD, CD38/CD157 has evolved more like an NADase, producing mainly ADP-ribose (ADPR) from NAD, with cADPR being a minor product. The acquisition of the NADase and the cADPR hydrolysis activities of CD38 make it an important signaling enzyme in regulating NAD and cADPR homeostasis (911). Genetic analysis, as well as the conservation of sequence and disulfide bonds among these enzymes, establish that they all evolved from a common ancestor (12). Little is known of why this conserved family of enzymes has evolved divergently in their catalytic metabolism of NAD and cADPR.Open in a separate windowFIGURE 1.Schemes of cADPR formation and mechanistic analogs for substrate and product. A, the cyclization reaction producing cADPR from NAD is catalyzed by both ADPRC and CD38. The structural difference between cADPR and N1-cIDPR lies at the 6-position of purine ring (6-NH for cADPR; 6-O for N1-cIDPR). B, an analog of the substrate NAD, N(2F-A)D, is enzymatically converted to 2F-ADPR by ADPRC instead of cyclized to c(2F-A)DPR. The formation of cADPR from NAD requires the intramolecular attack of the reaction intermediate by the adenine N1 atom. The addition of a fluorine atom on the adjacent C2 atom of adenine prevents the cyclization from occurring. C, ara-2′F-NAD and ribo-2′F-NAD are analogs of NAD that inhibit the cyclization reaction by producing covalent adducts during the catalysis by CD38. Both analogs differ only in the orientation of their fluorine atoms at the 2′-position of the adenine ribose.ADPRC, however, is not solely a cyclase because it can also catalyze the hydrolysis of NMN into ribose-5-phosphate and nicotinamide (13, 14). The catalytic outcome of this novel enzyme is thus determined not by the enzyme alone but also by the specific interactions between the active site and a particular substrate. Consistently, using an NAD analog, N(2F-A)D, as substrate, Zhang et al. (15) showed that the hydrolase activity of ADPRC can be dominantly revealed, whereas its cyclase activity is suppressed beyond detection (Fig. 1B). Likewise, human CD38 can be converted to a ADPRC-like enzyme by mutation of a single residue, Glu-146, at the active site (16). In this study, we determined the structural determinants critical for the catalytic characteristics of ADPRC and CD38 by comparing the crystal structures of the complexes of ADPRC and CD38 bound with various catalytically revealing substrates and products (Fig. 1, A–C). The results identify residues Phe-174 in the cyclase and Thr-221 in CD38 as the main determinants for the cyclase and hydrolysis activities of the enzymes. All together, these structures provide insights into the structural requirements for functional evolution from a cyclase to a hydrolase.  相似文献   

11.
NAD+ is mainly synthesized in human cells via the “salvage” pathways starting from nicotinamide, nicotinic acid, or nicotinamide riboside (NR). The inhibition with FK866 of the enzyme nicotinamide phosphoribosyltransferase (NAMPT), catalyzing the first reaction in the “salvage” pathway from nicotinamide, showed potent antitumor activity in several preclinical models of solid and hematologic cancers. In the clinical studies performed with FK866, however, no tumor remission was observed. Here we demonstrate that low micromolar concentrations of extracellular NAD+ or NAD+ precursors, nicotinamide mononucleotide (NMN) and NR, can reverse the FK866-induced cell death, this representing a plausible explanation for the failure of NAMPT inhibition as an anti-cancer therapy. NMN is a substrate of both ectoenzymes CD38 and CD73, with generation of NAM and NR, respectively. In this study, we investigated the roles of CD38 and CD73 in providing ectocellular NAD+ precursors for NAD+ biosynthesis and in modulating cell susceptibility to FK866. By specifically silencing or overexpressing CD38 and CD73, we demonstrated that endogenous CD73 enables, whereas CD38 impairs, the conversion of extracellular NMN to NR as a precursor for intracellular NAD+ biosynthesis in human cells. Moreover, cell viability in FK866-treated cells supplemented with extracellular NMN was strongly reduced in tumor cells, upon pharmacological inhibition or specific down-regulation of CD73. Thus, our study suggests that genetic or pharmacologic interventions interfering with CD73 activity may prove useful to increase cancer cell sensitivity to NAMPT inhibitors.  相似文献   

12.
Activation of NAD(P)H oxidase has been reported to produce superoxide (O2??) extracellularly as an autocrine/paracrine regulator or intracellularly as a signaling messenger in a variety of mammalian cells. However, it remains unknown how the activity of NAD(P)H oxidase is regulated in arterial myocytes. Recently, CD38-associated ADP-ribosylcyclase has been reported to use an NAD(P)H oxidase product, NAD+ or NADP+, to produce cyclic ADP-ribose (cADPR) or nicotinic acid adenine dinucleotide phosphate, which mediates intracellular Ca2 + signaling. This study was designed to test a hypothesis that the CD38/cADPR pathway as a downstream event exerts feedback regulatory action on the NAD(P)H oxidase activity in production of extra- or intracellular O2?? in mouse coronary arterial myocytes (CAMs). By fluorescence microscopic imaging, we simultaneously monitored extra- and intracellular O2?? production in wild-type (CD38+/+) and CD38 knockout (CD38?/?) CAMs in response to oxotremorine (OXO), a muscarinic type 1 receptor agonist. It was found that CD38 deficiency prevented OXO-induced intracellular but not extracellular O2?? production in CAMs. Consistently, the OXO-induced intracellular O2?? production was markedly inhibited by CD38 shRNA or the CD38 inhibitor nicotinamide in CD38+/+ CAMs. Further, Nox4 siRNA inhibited OXO-induced intracellular but not extracellular O2?? production, whereas Nox1 siRNA attenuated both intracellular and extracellular O2?? production in CD38+/+ CAMs. Direct delivery of exogenous cADPR into CAMs markedly elevated intracellular Ca2 + and O2?? production in CD38?/? CAMs. Functionally, CD38 deficiency or Nox1 siRNA and Nox4 siRNA prevented OXO-induced contraction in isolated perfused coronary arteries in CD38 WT mice. These results provide direct evidence that the CD38/cADPR pathway is an important controller of Nox4-mediated intracellular O2?? production and that CD38-dependent intracellular O2?? production is augmented in an autocrine manner by CD38-independent Nox1-derived extracellular O2?? production in CAMs.  相似文献   

13.
Nicotinamide adenine dinucleotide (NAD+) is crucial for cell energy metabolism and many signalling processes. Recently, we proved the role of ecto-enzymes in controlling adenine nucleotide–dependent pathways during calcific aortic valve disease (CAVD). This study aimed to investigate extracellular hydrolysis of NAD+ and mononucleotide nicotinamide (NMN) in aortic valves and aorta fragments of CAVD patients and on the inner aortic surface of ecto-5′-nucleotidase knockout mice (CD73−/−). Human non-stenotic valves (n = 10) actively converted NAD+ and NMN via both CD73 and NAD+-glycohydrolase (CD38) according to our analysis with RP-HPLC and immunofluorescence. In stenotic valves (n = 50), due to reduced CD73 activity, NAD+ was degraded predominantly by CD38 and additionally by ALP and eNPP1. CAVD patients had significantly higher hydrolytic rates of NAD+ (0.81 ± 0.07 vs 0.56 ± 0.10) and NMN (1.12 ± 0.10 vs 0.71 ± 0.08 nmol/min/cm2) compared with controls. CD38 was also primarily engaged in human vascular NAD+ metabolism. Studies using specific ecto-enzyme inhibitors and CD73−/− mice confirmed that CD73 is not the only enzyme involved in NAD+ and NMN hydrolysis and that CD38 had a significant contribution to these pathways. Modifications of extracellular NAD+ and NMN metabolism in aortic valve cells may be particularly important in valve pathology and could be a potential therapeutic target.  相似文献   

14.
Sterile Alpha and Toll Interleukin Receptor Motif–containing protein 1 (SARM1) is a key therapeutic target for diseases that exhibit Wallerian–like degeneration; Wallerian degeneration is characterized by degeneration of the axon distal to the site of injury. These diseases include traumatic brain injury, peripheral neuropathy, and neurodegenerative diseases. SARM1 promotes neurodegeneration by catalyzing the hydrolysis of NAD+ to form a mixture of ADPR and cADPR. Notably, SARM1 knockdown prevents degeneration, indicating that SARM1 inhibitors will likely be efficacious in treating these diseases. Consistent with this hypothesis is the observation that NAD+ supplementation is axoprotective. To identify compounds that block the NAD+ hydrolase activity of SARM1, we developed and performed a high–throughput screen (HTS). This HTS assay exploits an NAD+ analog, etheno–NAD+ (ENAD) that fluoresces upon cleavage of the nicotinamide moiety. From this screen, we identified berberine chloride and zinc chloride as the first noncompetitive inhibitors of SARM1. Though modest in potency, the noncompetitive mode of inhibition, suggests the presence of an allosteric binding pocket on SARM1 that can be targeted for future therapeutic development. Additionally, zinc inhibition and site–directed mutagenesis reveals that cysteines 629 and 635 are critical for SARM1 catalysis, highlighting these sites for the design of inhibitors targeting SARM1.  相似文献   

15.
Objectives: Mitochondrial oxidative stress is involved in the pathogenesis of diabetic kidney disease. The objective of our study is to identify the mechanisms of renal mitochondrial oxidative stress, focusing on Sirt3, which is nicotinamide adenine dinucleotide (NAD+; oxidized NAD)-dependent deacetylase in mitochondria.

Methods: Renal mitochondrial oxidative stress and Sirt3 activity, using Zucker diabetic fatty rats (ZDFRs) and cultured proximal tubular cells under high-glucose condition were evaluated.

Results: At 28 weeks of age, ZDFRs exhibited the increased urinary albumin/liver-type fatty acid-binding protein (L-FABP)/8-hydroxy-2'-deoxyguanosine (8-OHdG) excretion, histological tubular cell damage, compared to non-diabetic Zucker Lean rats. In renal mitochondria, acetylated isocitrate dehydrogenase2 (IDH2) and superoxide dismutase2 (SOD2), accompanied with mitochondrial oxidative stress and mitochondrial morphologic alterations, were increased in ZDFRs, indicating inactivation of Sirt3. Additionally, expression of the NAD-degrading enzyme, CD38, was increased, and the NAD+/NADH (reduced NAD) ratio was reduced in the renal cortex of ZDFRs. High-glucose stimulation in cultured proximal tubular cells also resulted in an increase in acetylated IDH2/SOD2, CD38 overexpression and a reduction in the NAD+/NADH ratio.

Conclusions: Enhancement of mitochondrial oxidative stress in the diabetic kidney was mediated by the reduction of Sirt3 activity. CD38 overexpression may be related to a reduction in the NAD+/NADH ratio in the diabetic kidney.  相似文献   


16.
Nicotinamide adenine dinucleotide (NAD+) is a key cofactor required for essential metabolic oxidation-reduction reactions. It also regulates various cellular activities, including gene expression, signaling, DNA repair and calcium homeostasis. Intracellular NAD+ levels are tightly regulated and often respond rapidly to nutritional and environmental changes. Numerous studies indicate that elevating NAD+ may be therapeutically beneficial in the context of numerous diseases. However, the role of NAD+ on skeletal muscle exercise performance is poorly understood. CD38, a multi-functional membrane receptor and enzyme, consumes NAD+ to generate products such as cyclic-ADP-ribose. CD38 knockout mice show elevated tissue and blood NAD+ level. Chronic feeding of high-fat, high-sucrose diet to wild type mice leads to exercise intolerance and reduced metabolic flexibility. Loss of CD38 by genetic mutation protects mice from diet-induced metabolic deficit. These animal model results suggest that elevation of tissue NAD+ through genetic ablation of CD38 can profoundly alter energy homeostasis in animals that are maintained on a calorically-excessive Western diet.  相似文献   

17.
The ectoenzyme CD38 catalyzes the production of cyclic ADP-ribose (cADPR) and ADP-ribose (ADPR) from its substrate, NAD(+). Both products of the CD38 enzyme reaction play important roles in signal transduction, as cADPR regulates calcium release from intracellular stores and ADPR controls cation entry through the plasma membrane channel TRPM2. We previously demonstrated that CD38 and the cADPR generated by CD38 regulate calcium signaling in leukocytes stimulated with some, but not all, chemokines and controls leukocyte migration to inflammatory sites. However, it is not known whether the other CD38 product, ADPR, also regulates leukocyte trafficking In this study we characterize 8-bromo (8Br)-ADPR, a novel compound that specifically inhibits ADPR-activated cation influx without affecting other key calcium release and entry pathways. Using 8Br-ADPR, we demonstrate that ADPR controls calcium influx and chemotaxis in mouse neutrophils and dendritic cells activated through chemokine receptors that rely on CD38 and cADPR for activity, including mouse FPR1, CXCR4, and CCR7. Furthermore, we show that the calcium and chemotactic responses of leukocytes are not dependent on poly-ADP-ribose polymerase 1 (PARP-1), another potential source of ADPR in some leukocytes. Finally, we demonstrate that NAD(+) analogues specifically block calcium influx and migration of chemokine-stimulated neutrophils without affecting PARP-1-dependent calcium responses. Collectively, these data identify ADPR as a new and important second messenger of mouse neutrophil and dendritic cell migration, suggest that CD38, rather than PARP-1, may be an important source of ADPR in these cells, and indicate that inhibitors of ADPR-gated calcium entry, such as 8Br-ADPR, have the potential to be used as anti-inflammatory agents.  相似文献   

18.
Mobilization of intracellular Ca2+ stores is involved in many diverse cell functions, including: cell proliferation; differentiation; fertilization; muscle contraction; secretion of neurotransmitters, hormones and enzymes; and lymphocyte activation and proliferation. Cyclic adenosine diphosphate ribose (cADPR) is an endogenous Ca2+ mobilizing nucleotide present in many cell types and species, from plants to animals. cADPR is formed by ADP-ribosyl cyclases from nicotinamide adenine dinucleotide. The main ADP-ribosyl cyclase in mammals is CD38, a multi-functional enzyme and a type II membrane protein. It has been shown that many extracellular stimuli can induce cADPR production that leads to calcium release or influx, establishing cADPR as a second messenger. cADPR has been linked to a wide variety of cellular processes, but the molecular mechanisms regarding cADPR signaling remain elusive. The aim of this review is to summarize the CD38/cADPR/Ca2+ signaling pathway, focusing on the recent advances involving the mechanism and physiological functions of cADPR-mediated Ca2+ mobilization.  相似文献   

19.
Abstract

Nicotinamide adenine dinucleotide (NAD) is a central metabolic cofactor by virtue of its redox capacity, and as such regulates a wealth of metabolic transformations. However, the identification of the longevity protein silent regulator 2 (Sir2), the founding member of the sirtuin protein family, as being NAD+-dependent reignited interest in this metabolite. The sirtuins (SIRT1-7 in mammals) utilize NAD+ to deacetylate proteins in different subcellular compartments with a variety of functions, but with a strong convergence on optimizing mitochondrial function. Since cellular NAD+ levels are limiting for sirtuin activity, boosting its levels is a powerful means to activate sirtuins as a potential therapy for mitochondrial, often age-related, diseases. Indeed, supplying excess precursors, or blocking its utilization by poly(ADP-ribose) polymerase (PARP) enzymes or CD38/CD157, boosts NAD+ levels, activates sirtuins and promotes healthy aging. Here, we discuss the current state of knowledge of NAD+ metabolism, primarily in relation to sirtuin function. We highlight how NAD+ levels change in diverse physiological conditions, and how this can be employed as a pharmacological strategy.  相似文献   

20.
HL-60, a human promyelocytic leukemia cell line, is induced to differentiate by retinoic acid to mature granulocytes. We have now found that after the addition of 1 μM retinoic acid to HL-60 cultures an increase in NAD+-glycohydrolase (NADase) activity is detected by 6 hr and after a 33-fold increase in activity reaches a plateau by 24 hr. Cycloheximide inhibits completely the retinoic acid-induced increase in NADase activity indicating that enzyme induction requires protein synthesis de, novo. An increase of NADase activity was found not only in HL-60 cells but also in two human monoblast cell lines (U-937 and THP-1) and fresh cells in primary culture from two patients with acute promyelocytic leukemia. An increase in synthesis de, novo of NADase does not appear to be obligatory for differentiation of HL-60 because there was no increase of NADase activity in HL-60 cells induced to differentiate with either dimethylsulfoxide, hypoxanthine, butyrate, or 1, 25-dihydroxycholecalciferol and there were marked increases in NADase activity at concentrations of retinoic acid having little or no effect on differentiation.  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