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.
In order to gain a better understanding of the role of ecto-NAD+ glycohydrolase, an enzyme predominantly associated with phagocytic cells, we have studied its fate in murine macrophages (splenic, resident peritoneal and Kupffer cells) during phagocytosis of opsonized on mannosylated latex beads. In parallel, we have also monitored nucleotide pyrophosphatase, another ecto-enzyme of macrophages. Phagosomes were isolated by flotation in a discontinuous sucrose gradient and the enzyme activities were determined with fluorometric methods. Low levels of NAD+ glycohydrolase and nucleotide pyrophosphatase could be measured associated with the phagosomal fractions, eg, respectively less than 4.5% and 10% in spleen macrophages. The phagosomal activities originate from the plasma membrane, ie they were latent and inactivation of ecto-NAD+ glycohydrolase with the diazonium salt of sulfanilic acid resulted in a marked decrease of this enzyme activity in the phagosomal fractions. Pre-labelling of the cell surface by [3H]-galactosylation indicated that NAD+ glycohydrolase is internalized to a lesser extent than an average surface-membrane unit. These results indicate that if ecto-NAD+ glycohydrolase of macrophages can be internalized to a limited extent during phagocytosis of opsonized or mannosylated latex beads, this enzyme appears to be predominantly excluded from the surface area involved in the uptake of such particles.  相似文献   

2.
《Bioorganic chemistry》1987,15(1):31-42
The use of NAD+ analogs lacking a carbonyl function at position C-3 of the pyridinium moiety allowed the manipulation of the kinetic mechanism of calf spleen NAD+ glycohydrolase so as to render the cleavage of the pyridinium-ribose bond rate limiting. The analogs used in this study are relatively poor substrates of the enzyme. They present an affinity for the active site which is independent of the nature of their substituent (Ki = 10 ± 2 μm), suggesting that the specificity of the NAD+ glycohydrolase reflects the dynamic steps occurring after the formation of the Michaelis complex. The maximal rates of hydrolysis of the NAD+ analogs are very sensitive to the pKa of the departing pyridine; a Brönsted plot (r = 0.99) gave a βtg = −0.90 (at 37°C). From this plot we could estimate that for NAD+, the specific interaction of the 3-carboxamide group with the active site contributed to the catalysis by decreasing the energy barrier by about 2 kcal mol−1. We have also studied the nonenzymatic hydrolysis of NAD+ and its analogs under conditions (pH-independent hydrolysis) which favor a unimolecular mechanism. In this case a linear Brönsted plot was also found (r = 0.99) with βtg = −1.11 (at 37°C). Our data indicate that NAD+ glycohydrolase catalyzes the chemical cleavage of the pyridinium-ribose bond, over a 103 rate difference, according to a single mechanism involving a late transition state in which the scissile bond is broken. The present study strongly supports our previous hypothesis (F. Schuber, P. Travo, and M. Pascal (1979) Bioorg. Chem. 8, 83) according to which NAD+ glycohydrolase catalyses unimolecular decomposition of its substrates with generation of an ADP-ribosyl oxocarbonium ion intermediate which must be stabilized by the active site of the enzyme.  相似文献   

3.
These studies show that both liver slices and macrophages carried out fibronectin concentration-dependent uptake of 125I-labeled gelatin-coated latex (test latex). Lack of phagocytosis of test latex by liver slices was shown directly by electron microscopy and indirectly by trypsin treatment, which caused the release of all test latex taken up in response to fibronectin. Inhibitors of phagocytosis did not alter this uptake. On the other hand, trypsin released only a portion of test latex from macrophages. Inhibitors of phagocytosis did not effect the released radioactive particles from macrophages but greatly reduced the trypsin-resistant radioactivity, taken as representing phagocytized particles. Opsonization of test latex with fibronectin did not require heparin but its association with liver slices occurred only in the presence of heparin. Macrophages, however, readily bound and internalized the opsonized test latex and heparin only potentiated these reactions. Gelatin competed with test latex for fibronectin for opsonization, but did not inhibit binding and phagocytosis of fibronectin-test latex complexes. Finally, soluble fibronectin-gelatin complexes did not compete for binding and phagocytosis of fibronectin-test latex complexes. Thus, fibronectin concentrated on the surface of latex is preferred for interaction with the fibronection receptor of macrophages. Gelatin, however, was not essential for this reaction, because fibronectin directly coupled to latex was also readily taken up.  相似文献   

4.
A 60- to 70-fold purification of an NAD+ glycohydrolase from the inner membrane of rat liver mitochondria to apparent homogeneity on sodium dodecyl sulfate (SDS)-polyacrylamide slab gel is described. The minimum molecular weight of the enzyme on polyacrylamide gels in the presence of SDS is around 62,000. The enzyme splits NAD+ to ADP-ribose and, presumably, nicotinamide. No phosphatase or phosphodiesterase activity is detected in the purified enzyme preparation. The enzyme shows high activity with NAD+ and moderate activity with NADP+ as substrates NAD(P)Hs are poor substrates. ATP and nicotinamide inhibit the enzyme. A possible participation of the enzyme in the mechanism of calcium release from rat liver mitochondria is discussed.  相似文献   

5.
The enzyme NAD+synthetase [deamido-NAD+: ammonia ligase (AMP-forming), EC 6.3.1.5] is used for the preparation of 2 μmol isotopically labelled [13N]NAD+, a radiopharmaceutical designed for positron emission tomography. To obtain a rapid and high yield synthesis of [13N]NAD+, the NAD+synthetase is immobilized on porous glass beads and packed in a column. The NAD+synthetase was obtained from Escherichia coli. Different strains were tested; the cell culture technique was optimized. A new, high yield purification was applied. A screening of different immobilization techniques was done. The selected immobilization method was further optimized to increase the enzymatic activity of the enzyme-loaded glass beads. The latter were packed into a glass column. The kinetic properties of this column were investigated and optimized.  相似文献   

6.
  • 1.1. Purified thyroidal NAD+ glycohydrolase has been subjected to the action of a number of group specific reagents in order to gain information concerning its mode of action.
  • 2.2. Modification of histidyl residues with diethylpyrocarbonate strongly suppresses the NAD+ glycohydrolase activity. Inactivation with this reagent can be reversed to some extent by subsequent treatment with hydroxylamine.
  • 3.3. NAD+ and ADP-ribose partially protect against inactivation with similar efficiencies.
  • 4.4. The incomplete reactivation with hydroxylamine after diethylpyrocarbonate treatment and the selective inactivation by 2,4-pentanedione indicates that apart from one or more essential histidyl residue(s) also lysyl residues are important for activity. NAD+ and to a smaller extent ADP-ribose again protect against inactivation by 2,4-pentanedione.
  • 5.5. The sensitivity of the enzyme towards N-ethyl-5-phenyl-isooxazolium-3'-sulfonate further points to the importance of carboxylate containing side chains.
  • 6.6. The mechanistic implications of these results are discussed.
  相似文献   

7.
Mitochondria from the parasitic helminth, Hymenolepis diminuta, catalyzed both NADPH:NAD+ and NADH:NADP+ transhydrogenase reactions which were demonstrable employing the appropriate acetylpyridine nucleotide derivative as the hydride ion acceptor. Thionicotinamide NAD+ would not serve as the oxidant in the former reaction. Under the assay conditions employed, neither reaction was energy linked, and the NADPH:NAD+ system was approximately five times more active than the NADH:NADP+ system. The NADH:NADP+ reaction was inhibited by phosphate and imidazole buffers, EDTA, and adenyl nucleotides, while the NADPH:NAD+ reaction was inhibited only slightly by imidazole and unaffected by EDTA and adenyl nucleotides. Enzyme coupling techniques revealed that both transhydrogenase systems functioned when the appropriate physiological pyridine nucleotide was the hydride ion acceptor. An NADH:NAD+ transhydrogenase system, which was unaffected by EDTA, or adenyl nucleotides, also was demonstrable in the mitochondria of H. diminuta. Saturation kinetics indicated that the NADH:NAD+ reaction was the product of an independent enzyme system. Mitochondria derived from another parasitic helminth, Ascaris lumbricoides, catalyzed only a single transhydrogenase reaction, i.e., the NADH:NAD+ activity. Transhydrogenase systems from both parasites were essentially membrane bound and localized on the inner mitochondrial membrane. Physiologically, the NADPH:NAD+ transhydrogenase of H. diminuta may serve to couple the intramitochondrial metabolism of malate (via an NADP linked “malic” enzyme) to the anaerobic NADH-dependent ATP-generating fumarate reductase system. In A. lumbricoides, where the intramitochondrial metabolism of malate depends on an NAD-linked “malic” enzyme which is localized primarily in the intermembrane space, the NADH:NAD+ transhydrogenase activity may serve physiologically in the translocation of hydride ions across the inner membrane to the anaerobic energy-generating fumarate reductase system.  相似文献   

8.
Functional morphodynamic behavior of differentiated macrophages is strongly controlled by actin cytoskeleton rearrangements, a process in which also metabolic cofactors ATP and NAD(H) (i.e. NAD+ and NADH) and NADP(H) (i.e. NADP+ and NADPH) play an essential role. Whereas the link to intracellular ATP availability has been studied extensively, much less is known about the relationship between actin cytoskeleton dynamics and intracellular redox state and NAD+-supply. Here, we focus on the role of nicotinamide phosphoribosyltransferase (NAMPT), found in extracellular form as a cytokine and growth factor, and in intracellular form as one of the key enzymes for the production of NAD+ in macrophages. Inhibition of NAD+ salvage synthesis by the NAMPT-specific drug FK866 caused a decrease in cytosolic NAD+ levels in RAW 264.7 and Maf-DKO macrophages and led to significant downregulation of the glycolytic flux without directly affecting cell viability, proliferation, ATP production capacity or mitochondrial respiratory activity. Concomitant with these differential metabolic changes, the capacity for phagocytic ingestion of particles and also substrate adhesion of macrophages were altered. Depletion of cytoplasmic NAD+ induced cell-morphological changes and impaired early adhesion in phagocytosis of zymosan particles as well as spreading performance. Restoration of NAD+ levels by NAD+, NMN, or NADP+ supplementation reversed the inhibitory effects of FK866. We conclude that direct coupling to local, actin-based, cytoskeletal dynamics is an important aspect of NAD+’s cytosolic role in the regulation of morphofunctional characteristics of macrophages.  相似文献   

9.
The Streptococcus pyogenes NAD+ glycohydrolase (SPN) is secreted from the bacterial cell and translocated into the host cell cytosol where it contributes to cell death. Recent studies suggest that SPN is evolving and has diverged into NAD+ glycohydrolase-inactive variants that correlate with tissue tropism. However, the role of SPN in both cytotoxicity and niche selection are unknown. To gain insight into the forces driving the adaptation of SPN, a detailed comparison of representative glycohydrolase activity-proficient and -deficient variants was conducted. Of a total 454 amino acids, the activity-deficient variants differed at only nine highly conserved positions. Exchanging residues between variants revealed that no one single residue could account for the inability of the deficient variants to cleave the glycosidic bond of β-NAD+ into nicotinamide and ADP-ribose; rather, reciprocal changes at 3 specific residues were required to both abolish activity of the proficient version and restore full activity to the deficient variant. Changing any combination of 1 or 2 residues resulted in intermediate activity. However, a change to any 1 residue resulted in a significant decrease in enzyme efficiency. A similar pattern involving multiple residues was observed for comparison with a second highly conserved activity-deficient variant class. Remarkably, despite differences in glycohydrolase activity, all versions of SPN were equally cytotoxic to cultured epithelial cells. These data indicate that the glycohydrolase activity of SPN may not be the only contribution the toxin has to the pathogenesis of S. pyogenes and that both versions of SPN play an important role during infection.  相似文献   

10.
After a 5-second exposure of illuminated bermudagrass (Cynodon dactylon L. var. `Coastal') leaves to 14CO2, 84% of the incorporated 14C was recovered as aspartate and malate. After transfer from 14CO2-air to 12CO2-air under continuous illumination, total radioactivity decreased in aspartate, increased in 3-phosphoglyceric acid and alanine, and remained relatively constant in malate. Carbon atom 1 of alanine was labeled predominantly, which was interpreted to indicate that alanine was derived from 3-phosphoglyceric acid. The activity of phosphoenolpyruvate carboxylase, alkaline pyrophosphatase, adenylate kinase, pyruvate-phosphate dikinase, and malic enzyme in bermudagrass leaf extracts was distinctly higher than those in fescue (Festuca arundinacea Schreb.), a reductive pentose phosphate cycle plant. Assays of malic enzyme activity indicated that the decarboxylation of malate was favored. Both malic enzyme and NADP+-specific malic dehydrogenase activity were low in bermudagrass compared to sugarcane (Saccharum officinarum L.). The activities of NAD+-specific malic dehydrogenase and acidic pyrophosphatase in leaf extracts were similar among the plant species examined, irrespective of the predominant cycle of photosynthesis. Ribulose-1, 5-diphosphate carboxylase in C4-dicarboxylic acid cycle plant leaf extracts was about 60%, on a chlorophyll basis, of that in reductive pentose phosphate cycle plants.  相似文献   

11.
Alveolar macrophages (AMs) can phagocytose unopsonized pathogens such as S. aureus via innate immune receptors, such as scavenger receptors (SRs). Cytoskeletal events and signaling pathways involved in phagocytosis of unopsonized bacteria likely govern the fate of ingested pathogens, but are poorly characterized. We have developed a high-throughput scanning cytometry-based assay to quantify phagocytosis of S. aureus by adherent human blood-derived AM-like macrophages in a 96-well microplate format. Differential fluorescent labeling of internalized vs. bound bacteria or beads allowed automated image analysis of collapsed confocal stack images acquired by scanning cytometry, and quantification of total particles bound and percent of particles internalized. We compared the effects of the classic SR blocker polyinosinic acid, the cytoskeletal inhibitors cytochalasin D and nocodazole, and the signaling inhibitors staurosporine, Gö 6976, JNK Inhibitor I and KN-93, on phagocytosis of a panel of live unopsonized S. aureus strains, (Wood, Seattle 1945 (ATCC 25923), and RN6390), as well as a commercial killed Wood strain, heat-killed Wood strain and latex beads. Our results revealed failure of the SR inhibitor polyinosinic acid to block binding of any live S. aureus strains, suggesting that SR-mediated uptake of a commercial killed fluorescent bacterial particle does not accurately model interaction with viable bacteria. We also observed heterogeneity in the effects of cytoskeletal and signaling inhibitors on internalization of different S. aureus strains. The data suggest that uptake of unopsonized live S. aureus by human macrophages is not mediated by SRs, and that the cellular mechanical and signaling processes that mediate S. aureus phagocytosis vary. The findings also demonstrate the potential utility of high-throughput scanning cytometry techniques to study phagocytosis of S. aureus and other organisms in greater detail.  相似文献   

12.
Three rat liver nucleoside(5′) diphosphosugar (NDP-sugar) or nucleoside(5′) diphosphoalcohol pyrophosphatases are described: two were previously identified in experiments measuring Mg2+-dependent ADP-ribose pyrophosphatase activity (Miró et al. (1989) FEBS Lett. 244, 123–126), and the other is a new, Mn2+-dependent ADP-ribose pyrophosphatase. They are resolved by ion-exchange chromatography, and differ by their substrate and cation specificities, KM values for ADP-ribose, pH-activity profiles, molecular weights and isoelectric points. The enzymes were tested for activity towards: reducing (ADP-ribose, IDP-ribose) and non-reducing NDP-sugars (ADP-glucose, ADP-mannose, GDP-mannose, UDP-mannose, UDP-glucose, UDP-xylose, CDP-glucose), CDP-alcohols (CDP-glycerol, CDP-ethanolamine, CDP-choline), dinucleotides (diadenosine pyrophosphate, NADH, NAD+, FAD), nucleoside(5′) mono- and diphosphates (AMP, CMP, GMP, ADP, CDP) and dTMP p-nitrophenyl ester. Since the enzymes have not been purified to homogeneity, more than three pyrophosphatases may be present, but the co-purification of activities, thermal co-inactivation, and inhibition experiments give support to: (i) an ADP-ribose pyrophosphatase highly specific for ADP(IDP)-ribose in the presence of Mg2+, but active also on non-reducing ADP-hexoses and dinucleotides (not on NAD+) when Mg2+ was replaced with Mn2+; (ii) a Mn2+-dependent pyrophosphatase active on ADP(IDP)-ribose, dinucleotides and CDP-alcohols; (iii) a rather unspecific pyrophosphatase that, with Mg2+, was active on AMP(IMP)-containing NDP-sugars and dinucleotides (not on NAD+), and with Mn2+, was also active on non-adenine NDP-sugars and CDP-alcohols. The enzymes differ from nucleotide pyrophosphatase/phosphodiesterase-I (NPPase/PDEaseI) by their substrate specificities and by their cytosolic location and solubility in the absence of detergents. Although NPPase/PDEaseI is much more active in rat liver, its known location in the non-cytoplasmic sides of plasma and endoplasmic reticulum membranes, together with the known cytoplasmic synthesis of NDP-sugars and CDP-alcohols, permit the speculation that the pyrophosphatases studied in this work may have a cellular role.  相似文献   

13.
We have tested the feasibility of targeting liposomes via interaction with specific ecto-enzymes, i.e., enzymes which have their active site oriented to the external surface of the cell. 3,4-Dimethylpyridine adenine dinucleotide, a competitive inhibitor of ecto-NAD+-glycohydrolase, was substituted at N6 with a hydrophilic spacer arm, functionalized with a sulfhydryl group, and covalently linked to preformed liposomes containing 4-(p-maleimidophenyl)butyryl phosphatidylethanolamine. We show that compared to control vesicles, the binding of the conjugated liposomes was greatly increased (up to 5-fold) to cells presenting ecto-NAD+-glycohydrolase activity (Swiss 3T3 fibroblasts, mouse peritoneal macrophages); in contrast, no specific binding was detected with hepatoma tissue culture cells, which lack this enzyme. Specific binding was found to depend on the ligand/lipid molar ratio of the vesicles and on the length of the arm. High concentrations of free 3,4-dimethylpyridine adenine dinucleotide virtually abolished the specific binding to cells of the targeted liposomes. Analysis of binding revealed that the ligand conjugated to the liposomes presented a functional affinity for 3T3 fibroblasts 15-fold superior to that of the free ligand.  相似文献   

14.
These studies show that both liver slices and macrophages carried out fibronectin concentration-dependent uptake of 125I-labeled gelatin-coated latex (test latex). Lack of phagocytosis of test latex by liver slices was shown directly by electron microscopy and indirectly by trypsin treatment, which caused the release of all test latex taken up in response to fibronectin. Inhibitors of phagocytosis did not alter this uptake. On the other hand, trypsin released only a portion of test latex from macrophages. Inhibitors of phagocytosis did not effect the released radioactive particles from macrophages but greatly reduced the trypsin-resistant radioactivity, taken as representing phagocytized particles. Opsonization of test latex with fibronectin did not require heparin but its association with liver slices occurred only in the presence of heparin. Macrophages, however, readily bound and internalized the opsonized test latex and heparin only potentiated these reactions. Gelatin competed with test latex for fibronectin for opsonization, but did not inhibit binding and phagocytosis of fibronectin-test latex complexes. Finally, soluble fibronectin-gelatin complexes did not compete for binding and phagocytosis of fibronectin-test latex complexes. Thus, fibronectin concentrated on the surface of latex is preferred for interaction with the fibronectin receptor of macrophages. Gelatin, however, was not essential for this reaction, because fibronectin directly coupled to latex was also readily taken up.  相似文献   

15.
In this article we compare the kinetic behavior toward pyridine nucleotides (NAD+, NADH) of NAD+-malic enzyme, pyruvate dehydrogenase, isocitrate dehydrogenase, α-ketoglutarate dehydrogenase, and glycine decarboxylase extracted from pea (Pisum sativum) leaf and potato (Solanum tuberosum) tuber mitochondria. NADH competitively inhibited all the studied dehydrogenases when NAD+ was the varied substrate. However, the NAD+-linked malic enzyme exhibited the weakest affinity for NAD+ and the lowest sensitivity for NADH. It is suggested that NAD+-linked malic enzyme, when fully activated, is able to raise the matricial NADH level up to the required concentration to fully engage the rotenone-resistant internal NADH-dehydrogenase, whose affinity for NADH is weaker than complex I.  相似文献   

16.
This study concerned the role of the sulfhydryl groups in urocanase of Pseudomonas putida. When p-chloromercuribenzoate was added to the enzyme, two sulfhydryl groups reacted at once with little inhibition; the enzyme slowly became inhibited while further sulfhydryls reacted. After the p-chloromercuribenzoate inhibition occurred, if a thiol was subsequently added, most of the original activity was recovered. As the incubation time with p-chloromercuribenzoate was increased, the thiol became less effective in reversing the inhibition. However, if NAD+ (10 μm) was added with the thiol, 60–90% of the initial activity was restored even after long p-chloromercuribenzoate incubations. Restoration of activity by NAD+ was concentration dependent and specific for NAD+. Radioactive NAD+ could be bound to urocanase. These results confirm the coenzyme role for NAD+ in urocanase. In urea, p-chloromercuribenzoate titration of urocanase measured 11.9 -SH groups per molecule. Sulfite-modified enzyme treated with p-chloromercuribenzoate and dialyzed was substantially photoactivated in the presence of a thiol; that is, NAD+ was not required to restore activity. From these results, it is proposed that this enzyme contains two reactive —SH groups and that an essential —SH group is involved in NAD+ binding. Forces present in the sulfite-modified enzyme prevent the release of the NAD+ in the presence of mercurials.  相似文献   

17.
Intrinsic fluorescence of polystyrene dissolved in organic solvents such as 1,2-dimethoxyethane was used to develop a sensitive method for the quantification of polystyrene latex beads. This method allows the assay of latex in the microgram range and is one order of magnitude more sensitive than the conventional spectrophotometric method. The fluorometric technique was used in the quantification of phagocytic latex particle uptake by macrophages and in the quantification of isolated phagosomal fractions.  相似文献   

18.
NAD+-dependent and NADP+-dependent glyceraldehyde-3-phosphate (G-3-P) dehydrogenases were isolated from Euglena gracilis and characterized as to their physical and chemical parameters. NAD+-G-3-P dehydrogenase was found to have a strong resemblance to similar enzymes from muscle tissue. It has a molecular weight of about 140,000, four subunits of identical size and charge, and a single species of NH2-terminal amino acid. Two sulfhydryl groups per subunit are present, one of which is directly involved in the catalytic activity and is rapidly titratable. The enzyme also exhibits the “half the sites reactivity” of sulfhydryl groups as defined by O. P. Malhotra and S. A. Bernhard ((1968) J. Biol. Chem. 243, 1243). The pH and temperature optima are also similar to those of the enzymes from muscle tissue, as are the reaction kinetics and the strict specificity for NAD+.NADP+-dependent G-3-P dehydrogenase is different in many respects. Its molecular weight is slightly lower (~136,000) than that of the NAD+ enzyme, though it also consists of four subunits. It has a higher affinity for the reverse reaction substrates, in line with its probable function in vivo in CO2 fixation. There is only one sulfhydryl group per subunit, and that is not involved in activity, suggesting a difference in reaction mechanisms between the two enzymes. The NADP+-dependent enzyme exhibits activation by ATP, whereas the NAD+-dependent enzyme is competitively inhibited by this nucleotide.The greatest difference observed is in the physical characteristics of the enzymes. NADP+-G-3-P dehydrogenase was highly hydrophobic. Its solubility in a 10% aqueous solution of p-dioxane was approximately four to five times that of the NAD+-enzyme. Isolation of the enzyme was accomplished by fractionation in 1,2-dimethoxyethane, which also stabilized the enzymatic activity, as did aqueous p-dioxane. The high axial ratio of the NADP+-enzyme (~9) coupled with its very low degree of hydration as well as the high degree of amidation of the dicarboxylic amino acids (>90%) indicates that the exterior of the enzyme molecule is probably hydrophobic in nature. This is in agreement with its in vivo hydrophobic environment in the chloroplast membrane and explains the lability of the enzyme once extracted into an aqueous environment as well as its stabilization in solvents.  相似文献   

19.
The hydrolysis of NAD+ analogs bearing different substituents in the pyridinium moiety, catalyzed by solubilized calf spleen NAD+ glycohydrolase, was studied. The enzyme was specific for analogs possessing a carbonyl function at position C-3. The observed maximal rates showed no simple dependence either on the leaving ability of the parent pyridines or on the observed binding energies. The analogs without carbonyl substituents were not found to be hydrolyzed under our experimental conditions; and, compared to the substrates, they all presented much higher affinities for the active site, which could not be related to specific interactions. These results indicate that the rate-limiting step of the NAD+ hydrolysis, which is the formation of an enzyme-ADP-ribosyl intermediate, is probably more complex than a simple chemical step, i.e., pyridinium-ribose bond breakage. At a molecular level we favor a catalytic mechanism which, through nonbonded interactions between the substrate and the active site of the enzyme, results in the destabilization of the pyridinium-ribose bond, i.e., unimolecular decomposition involving an oxocarbonium ion intermediate.  相似文献   

20.
ADP-ribosyl cyclase synthesizes the secondary messenger cyclic ADP-ribose from NAD+. Diffraction quality crystals of the enzyme from ovotestes of Aplysia californica have been obtained. Crystallographic analysis of this enzyme will yield insight into the mode of binding of the novel cyclic nucleotide and the mechanism by which NAD+ is cyclized.  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