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 39 毫秒
1.
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   

2.
Bovine CD38/NAD+ glycohydrolase catalyzes the hydrolysis of NAD+ to nicotinamide and ADP-ribose and the formation of cyclic ADP-ribose via a stepwise reaction mechanism. Our recent crystallographic study of its Michaelis complex and covalently-trapped intermediates provided insights into the modalities of substrate binding and the molecular mechanism of bCD38. The aim of the present work was to determine the precise role of key conserved active site residues (Trp118, Glu138, Asp147, Trp181 and Glu218) by focusing mainly on the cleavage of the nicotinamide–ribosyl bond. We analyzed the kinetic parameters of mutants of these residues which reside within the bCD38 subdomain in the vicinity of the scissile bond of bound NAD+. To address the reaction mechanism we also performed chemical rescue experiments with neutral (methanol) and ionic (azide, formate) nucleophiles. The crucial role of Glu218, which orients the substrate for cleavage by interacting with the N-ribosyl 2′-OH group of NAD+, was highlighted. This contribution to catalysis accounts for almost half of the reaction energy barrier. Other contributions can be ascribed notably to Glu138 and Asp147 via ground-state destabilization and desolvation in the vicinity of the scissile bond. Key interactions with Trp118 and Trp181 were also proven to stabilize the ribooxocarbenium ion-like transition state. Altogether we propose that, as an alternative to a covalent acylal reaction intermediate with Glu218, catalysis by bCD38 proceeds through the formation of a discrete and transient ribooxocarbenium intermediate which is stabilized within the active site mostly by electrostatic interactions.  相似文献   

3.
The kinetic α-secondary deuterium isotope effect, kHkD, for the pH-independent hydrolysis of nicotinamide riboside, yielding nicotinamide and ribose, in water at 25 ° is 1.14, establishing that this reaction proceeds with unimolecular substrate decomposition to yield a carboxonium ion, or related species, in the rate-determining step. Surprisingly, the corresponding isotope effect for the base-catalyzed decomposition of the same substrate is 1.12, a value indicating considerable sp2 character at the Cl′ position in the transition state for this reaction. A similar result, kHkD = 1.15, was obtained for base-catalyzed hydrolysis of NAD+. The kinetic alpha deuterium isotope effect for the pig brain NAD glycohydrolasecatalyzed hydrolysis of nicotinamide riboside is 1.08. This value suggests that CN bond cleavage to form an intermediate carboxonium ion, or structurally related species, is at least partially rate-determining. In contrast, the corresponding value for the hydrolysis of this substrate catalyzed by Escherichia coli nicotinamide ribonucleotide glycohydrolase is very near unity, a result consistent with several interpretations including a rate-determining enzyme isomerization reaction.  相似文献   

4.
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 or mannosylated latex beads. In parallel, we have also monitored nucleotide pyrophosphatase, another ectoenzyme 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   

5.
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   

6.
The validity of the “charge-relay” system in serine esterases was examined by use of the general base-catalyzed hydrolysis of ethyl chloroacetate (I) as a model system. The general base catalytic rate for 2-benzimidazoleacetic acid (II) exhibited an eightfold positive deviation from the Brønsted plot including benzimidazole, imidazole, N-methylimidazole (V), and acetate ion, though in nucleophilic catalysis of the hydrolysis of p-nitrophenyl acetate, the point for II conformed to the Brønsted relationship together with imidazole and benzimidazole derivatives. The positive deviation of II from the Brønsted plot for the general base-catalysis was attributed to the cooperativity of the carboxyl group of II, the imidazolyl group of II, and the hydroxyl group of water. The present result provides support for the “charge-relay” system. Furthermore, the (essentially) total loss of the enzymatic activity due to N-3 methylation of histidine-57 in α-chymotrypsin is discussed in comparison to the general base-catalysis of V in the hydrolysis of I, which is also favorable for the “charge-relay” system.  相似文献   

7.
All the equilibrium conformations of 34 analogues of acetylcholine (ACh) with the general formula R-C(O)O-Alk-N+(CH3)3 are calculated by the method of molecular mechanics. In the series R-C(O)O-(CH2)2-N+(CH3)3, a reliable correlation is found between the molecular volume of the substrate and the rate of its hydrolysis by acetylcholinesterase (AChE); the absence of such a correlation is demonstrated for butyryl-cholinesterase (BChE). Theoretical conformational analysis confirms that the completely extended tt conformation of ACh is productive for the hydrolysis by AChE, which agrees with the results of X-ray analysis of AChE. AChE is shown to hydrolyze only those substrates that form equilibrium conformers compatible in the mutual arrangement of trimethylammonium group, carbonyl carbon, and carbonyl oxygen with the tt conformation of ACh; in this case, the rate of substrate hydrolysis depends on the total population of these conformers. A reliable correlation was found between the population of the semifolded (tg?) conformation of the choline moiety of substrate molecules and rate of their BChE hydrolysis. In a series of CH3-C(O)O-Alk-N+(CH3)3, the rate of BChE hydrolysis is demonstrated to depend on the total population of conformations compatible in the mutual arrangement of functionally important atoms with the tg? conformation of ACh. The tg? conformation of ACh is concluded to be productive for BChE hydrolysis. Similar orientations of the substrate molecules relative to the catalytic triads of both AChE and BChE are proven to coincide upon the substrate productive sorption in their active sites. It is hypothesized that the sorption stage is rate-limiting in cholinesterase hydrolysis and the enzyme hydrolyzes the ACh molecule in its energetically favorable conformation.  相似文献   

8.
Robert F. Anderson 《BBA》1980,590(2):277-281
The one-electron reduction potential of NAD+ has been determined using pulse radiolysis to study electron-transfer equilibria between it and a low potential bipyridylium compound. The determined value E17 (NAD+/NAD.) = ?922 ± 8 mV (NHE scale) is used to calculate E27 (NAD./NADH) = +282 mV. E17 for 1-methylnicotinamide, E17 (MeN+/MeN.) = ?918 ± 7 mV.  相似文献   

9.
  • 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   

10.
Glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides utilizes either NAD+ or NADP+ as coenzyme. Kinetic studies showed that NAD+ and NADP+ interact with different enzyme forms (Olive, C., Geroch, M. E., and Levy, H. R. (1971) J. Biol. Chem.246, 2047–2057). In the present study the techniques of fluorescence quenching and fluorescence enhancement were used to investigate the interaction between Leuconostoc mesenteroides glucose-6-phosphate dehydrogenase and coenzymes. In addition, kinetic studies were performed to examine interaction between the enzyme and various coenzyme analogs. The maximum quenching of protein fluorescence is 5% for NADP+ and 50% for NAD+. The dissociation constant for NADP+, determined from fluorescence quenching measurements, is 3 μm, which is similar to the previously determined Km of 5.7 μm and Ki of 5 μm. The dissociation constant for NAD+ is 2.5 mm, which is 24 times larger than the previously determined Km of 0.106 mm. Glucose 1-phosphate, a substrate-competitive inhibitor, lowers the dissociation constant and maximum fluorescence quenching for NAD+ but not for NADP+. This suggests that glucose 6-phosphate may act similarly and thus play a role in enabling the enzyme to utilize NAD+ under physiological conditions. When NADPH binds to the enzyme its fluorescence is enhanced 2.3-fold. The enzyme was titrated with NADPH in the absence and presence of NAD+; binding of these two coenzymes is competitive. The dissociation constant for NADPH from these measurements is 24 μm; the previously determined Ki is 37.6 μm. The dissociation constant for NAD′ is 2.8 mm, in satisfactory agreement with the value obtained from protein fluorescence quenching measurements. Various compounds which resemble either the adenosine or the nicotinamide portion of the coenzyme structure are coenzyme-competitive inhibitors; 2′,5′-ADP, the most inhibitory analog tested, gives NADP+-competitive and NAD+-noncompetitive inhibition, consistent with the kinetic mechanism previously proposed. By using pairs of coenzyme-competitive inhibitors it was shown in kinetic studies that the two portions of the NAD+ structure cannot be accommodated on the enzyme simultaneously unies they are covalently linked. Fluorescence studies showed that there are both “buried” and “exposed” tryptophan residues in the enzyme structure.  相似文献   

11.
《Inorganica chimica acta》1988,149(1):151-154
The extraction equilibrium of the hydronium-uranium(VI)-dicyclohexano-24-crown-8 complex was carried out in the crown ether1,2-dichloroethaneHCl aqueous solution system at different temperatures. The extraction complex has the overall composition (L)2·(H3O+·χH2O)2·UO2Cl42− (L = dicyclohexano-24-crown-8). The values of the extraction equilibrium constants (Kex) increase steadily with a decrease in temperature: 13.5 (298 K), 7.96 (301 K), 4.20 (303 K) and 2.07 (305 K). A plot of log Kex against 1/T shows a straight line. The value of the enthalpy change, ΔH°, was calculated from the slope and equals −212 kJ mol−1. The value of the entropy change, ΔS°, was calculated from ΔH° and Kex and equals −690 J K−1 mol−1, whereas ΔG° = −6.45 kJ mol−1. Comparing these thermodynamic parameters with those of the dicyclohexano-18-crown-6 isomer A [1] (ΔS° = −314 J K−1 mol−1, ΔH° = −101 kJ mol−1 and ΔG° = −8.37 kJ mol−1), it can be seen that ΔH° and ΔS° are more negative for the former than for the latter, and both are enthalpy-stabilized complexes. The molecular structure of the complex has the feature that there are two H5O2+ ions in it, in contrast to the H3O+ ions in the dicyclohexano-18-crown-6 isomer A complex [1]. Each of the H5O2+ ions is held in the crown ether cavity by four hydrogen bonds. The H5O2+ ion has a central bond. The uranium atom forms UO2Cl42− as a counterion away from the crown ether. The formation of this complex is in good agreement with more negative entropy change and less negative free energy change, as mentioned above.  相似文献   

12.
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   

13.
Addition of NAD+ to purified potato (Solanum tuberosum L.) mitochondria respiring α-ketoglutarate and malate in the presence of the electron transport inhibitor rotenone, stimulated O2 uptake. This stimulation was prevented by incubating mitochondria with N-4-azido-2-nitrophenyl-aminobutyryl-NAD+ (NAP4-NAD+), an inhibitor of NAD+ uptake, but not by 1 mm EGTA, an inhibitor of external NADH oxidation. NAD+-stimulated malate-cytochrome c reductase activity, and reduction of added NAD+ by intact mitochondria, could be duplicated by rupturing the mitochondria and adding a small quantity to the cuvette. The extent of external NAD+ reduction was correlated with the amount of extra mitochondrial malate dehydrogenase present. Malate oxidation by potato mitochondria depleted of endogenous NAD+ by storing on ice for 72 hours, was completely dependent on added NAD+, and the effect of NAD+ on these mitochondria was prevented by incubating them with NAP4-NAD+. External NAD+ reduction by these mitochondria was not affected by NAP4-NAD+. We conclude that all effects of exogenous NAD+ on plant mitochondrial respiration can be attributed to net uptake of the NAD+ into the matrix space.  相似文献   

14.
Phosphite dehydrogenase (PTDH) catalyzes the NAD+-dependent oxidation of phosphite to phosphate. This reaction requires the deprotonation of a water nucleophile for attack on phosphite. A crystal structure was recently solved that identified Arg301 as a potential base given its proximity and orientation to the substrates and a water molecule within the active site. Mutants of this residue showed its importance for efficient catalysis, with about a 100-fold loss in k cat and substantially increased K m,phosphite for the Ala mutant (R301A). The 2.35 Å resolution crystal structure of the R301A mutant with NAD+ bound shows that removal of the guanidine group renders the active site solvent exposed, suggesting the possibility of chemical rescue of activity. We show that the catalytic activity of this mutant is restored to near wild-type levels by the addition of exogenous guanidinium analogues; Brønsted analysis of the rates of chemical rescue suggests that protonation of the rescue reagent is complete in the transition state of the rate-limiting step. Kinetic isotope effects on the reaction in the presence of rescue agents show that hydride transfer remains at least partially rate-limiting, and inhibition experiments show that K i of sulfite with R301A is ∼400-fold increased compared to the parent enzyme, similar to the increase in K m for phosphite in this mutant. The results of our experiments indicate that Arg301 plays an important role in phosphite binding as well as catalysis, but that it is not likely to act as an active site base.  相似文献   

15.
The NAD+-dependent isocitrate dehydrogenase from etiolated pea (Pisum sativum L.) mitochondria was purified more than 200-fold by dye-ligand binding on Matrix Gel Blue A and gel filtration on Superose 6. The enzyme was stabilized during purification by the inclusion of 20% glycerol. In crude matrix extracts, the enzyme activity eluted from Superose 6 with apparent molecular masses of 1400 ± 200, 690 ± 90, and 300 ± 50 kD. During subsequent purification steps the larger molecular mass species disappeared and an additional peak at 94 ± 16 kD was evident. The monomer for the enzyme was tentatively identified at 47 kD by sodium dodecyl-polyacrylamide gel electrophoresis. The NADP+-specific isocitrate dehydrogenase activity from mitochondria eluted from Superose 6 at 80 ± 10 kD. About half of the NAD+ and NADP+-specific enzymes remained bound to the mitochondrial membranes and was not removed by washing. The NAD+-dependent isocitrate dehydrogenase showed sigmodial kinetics in response to isocitrate (S0.5 = 0.3 mm). When the enzyme was aged at 4°C or frozen, the isocitrate response showed less allosterism, but this was partially reversed by the addition of citrate to the reaction medium. The NAD+ isocitrate dehydrogenase showed standard Michaelis-Menten kinetics toward NAD+ (Km = 0.2 mm). NADH was a competitive inhibitor (Ki = 0.2 mm) and, unexpectedly, NADPH was a noncompetitive inhibitor (Ki = 0.3 mm). The regulation by NADPH may provide a mechanism for coordination of pyridine nucleotide pools in the mitochondria.  相似文献   

16.
17.
Sirtuins are NAD+ dependent deacetylases that modulate various essential cellular functions. Development of peptide based inhibitors of Sir2s would prove useful both as pharmaceutical agents and as effectors by which downstream cellular alterations can be monitored. Click chemistry that utilizes Huisgen’s 1,3-dipolar cycloaddition permits attachment of novel modifications onto the side chain of lysine. Herein, we report the synthesis of peptide analogues prepared using click reactions on -propargyloxycarbonyl protected lysine residues and their characterization as inhibitors of Plasmodium falciparum Sir2 activity. The peptide based inhibitors exhibited parabolic competitive inhibition with respect to acetylated-peptide substrate and parabolic non-competitive inhibition with NAD+ supporting the formation of EI2 and E·NAD+·I2 complexes. Cross-competition inhibition analysis with the non-competitive inhibitor nicotinamide (NAM) ruled out the possibility of the NAM-binding site being the second inhibitor binding site, suggesting the presence of a unique alternate pocket accommodating the inhibitor. One of these compounds was also found to be a potent inhibitor of the intraerythrocytic growth of P. falciparum with 50% inhibitory concentration in the micromolar range.  相似文献   

18.
Homogeneous preparations of l-threonine dehydrogenase (l-threonine: NAD+ oxidoreductase, EC 1.1.1.103) from Escherichia coli K-12, after having been dialyzed against buffers containing Chelex-100 resin, have a basal level of activity of 10–20 units/mg. Added Cd2+ stimulates dehydrogenase activity approx. 10-fold; this activation is concentration-dependent and is saturable with an activation Kd = 0.9 μM. Full activation by Cd2+ is obtained in the absence of added thiols. The pH-activity profile of the Cd2+-activated enzyme conforms to a theoretical curve for one-proton ionization with a pKa = 7.85. Mn2+, the only other activating metal ion, competes with Cd2+ for the same binding site. Km values forl-threonine and NAD+ as well as the Vmax for ‘demetallized’, Cd2+-activated, and Mn2+-activated threonine dehydrogenase were determined and compared.  相似文献   

19.
The subcellular distribution of NADP+ and NAD+-dependent glucose-6-phosphate and galactose-6-phosphate dehydrogenases were studied in rat liver, heart, brain, and chick brain. Only liver particulate fractions oxidized glucose-6-phosphate and galactose-6-phosphate with either NADP+ or NAD+ as cofactor. While all of the tissues examined had NADP+-dependent glucose-6-phosphate dehydrogenase activity, only rat liver and rat brain soluble fractions had NADP+-dependent galactose-6-phosphate dehydrogenase activity. Rat liver microsomal and rat brain soluble galactose-6-phosphate dehydrogenase activities were kinetically different (Km's 0.5 mm and 10 mm, respectively, for galactose-6-phosphate), although their reaction products were both 6-phosphogalactonate. Rat brain subcellular fractions did not oxidize 6-phosphogalactonate with either NADP+ or NAD+ cofactors but phosphatase activities hydrolyzing 6-phosphogalactonate, galactose-6-phosphate and galactose-1-phosphate were found in crude brain homogenates. In addition, galactose-6-phosphate and 6-phosphogalactonate were tested as inhibitors of various enzymes, with largely negative results, except that 6-phosphogalactonate was a competitive inhibitor (Ki = 0.5 mM) of rat brain 6-phosphogluconate dehydrogenase.  相似文献   

20.
A mechanism by which intact potato (Solanum tuberosum) mitochondria may regulate the matrix NAD content was studied in vitro. If mitochondria were incubated with NAD+ at 25°C in 0.3 molar mannitol, 10 millimolar phosphate buffer (pH 7.4), 5 millimolar MgCl2, and 5 millimolar α-ketoglutarate, the NAD pool size increased with time. In the presence of uncouplers, net uptake was not only inhibited, but NAD+ efflux was observed instead. Furthermore, the rate of NAD+ accumulation in the matrix space was strongly inhibited by the analog N-4-azido-2-nitrophenyl-4-aminobutyryl-3′-NAD+. When suspended in a medium that avoided rupture of the outer membrane, intact purified mitochondria progressively lost their NAD+ content. This led to a slow decrease of NAD+-linked substrates oxidation by isolated mitochondria The rate of NAD+ efflux from the matrix space was strongly temperature dependent and was inhibited by the analog inhibitor of NAD+ transport indicating that a carrier was required for net flux in either direction. It is proposed that uptake and efflux operate to regulate the total matrix NAD pool size.  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