首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Choleragen catalyzed the hydrolysis of NAD to ADP-ribose and nicotinamide; nicotinamide production was dramatically increased by L-arginine methyl ester and to a lesser extent by D- or L-arginine, but not by other basic amino acids. Guanidine was also effective. Nicotinamide formation in the presence of L-arginine methyl ester was greatest under conditions previously shown to accelerate the hydrolysis of NAD by choleragen (Moss, J., Manganiello, V. C., and Vaughan, M. (1976) Proc. Natl. Acad. Sci. U.S.A. 73, 4424-4427). After incubation of [adenine-U14C]NAD and L[3H]arginine with coleragen, a product was isolated by thin layer chromatography that contained adenine and arginine in a 1:1 ratio and has been tentatively identified as ADP-ribose-L-arginine. Parallel experiments with [carbonyl-14C]NAD have demonstrated that formation of the ADP-ribosyl-L-arginine derivative was associated with the production of [carbonyl-14C]nicotinamide. As guanidine itself was active and D- and L-arginine was equally effective in promoting nicotinamide production, whereas citrulline, which possesses a ureido rather than a guanidino function, was inactive, it seems probable that the guanidino group rather than the alpha-amino moiety participated in the linkage to ADP-ribose. Based on the assumption that the ADP-ribosylation of L-arginine by choleragen is a model for the NAD-dependent activation of adenylate cyclase by choleragen, it is proposed that the active A protomer of choleragen catalyzes the ADP-ribosylation of an arginine, or related amino acid residue in a protein, which is the cyclase itself or is critical to its activation by choleragen.  相似文献   

2.
Choleragen (cholera toxin) activates adenylate cyclase by catalyzing ADP-ribosylation of Gs alpha, the stimulatory guanine nucleotide-binding protein. It was recently found (Tsai, S.-C., Noda, M., Adamik, R., Moss, J., and Vaughan, M. (1987) Proc. Natl. Acad. Sci. U. S. A. 84, 5139-5142) that a bovine brain membrane protein known as ADP-ribosylation factor or ARF, which enhances ADP-ribosylation of Gs alpha, also increases the GTP-dependent NAD:arginine and NAD:protein ADP-ribosyltransferase, NAD glycohydrolase, and auto-ADP-ribosylation activities of choleragen. We report here the purification and characterization of two soluble proteins from bovine brain that similarly enhance the Gs alpha-dependent and independent ADP-ribose transfer reactions catalyzed by toxin. Like membrane ARF, both soluble factors are 19-kDA proteins dependent on GTP or GTP analogues for activity. Maximal ARF effects were observed at a molar ratio of less than 2:1, ARF/toxin A subunit. Dimyristoyl phosphatidylcholine was necessary for optimal ADP-ribosylation of Gs alpha but inhibited auto-ADP-ribosylation of the choleragen A1 subunit and NAD:agmatine ADP-ribosyltransferase activity. It appears that the soluble factors directly activate choleragen in a GTP-dependent fashion. The relationships of the ARF proteins to the ras oncogene products and to the family of guanine nucleotide-binding regulatory proteins that includes Gs alpha remains to be determined.  相似文献   

3.
Linear terminal alkane-diols have been shown to function as more efficient substrates of the transglycosylational activity of NAD+ glycohydrolase (NADase) than the corresponding respective 1-alkanols. A series of eight alkane-diols from ethane-1,2- to nonane-1,9-diols underwent an O-ADP-ribosylation in the incubation reaction with NAD/NADase to provide the corresponding ribosylated products. The structural properties of these products were characterized by 1H NMR and MS spectrometries. No substantial double ADP-ribosylation of the two hydroxy functions was observed which was initially expected in the diols of higher carbon number.  相似文献   

4.
The bacterial toxins, choleragen and pertussis toxin, inhibit the light-stimulated GTPase activity of bovine retinal rod outer segments by catalysing the ADP-ribosylation of the alpha-subunit (T alpha) of transducin [Abood, Hurley, Pappone, Bourne & Stryer (1982) J. Biol. Chem. 257, 10540-10543; Van Dop, Yamanaka, Steinberg, Sekura, Manclark, Stryer & Bourne (1984) J. Biol. Chem. 259, 23-26]. Incubation of retinal rod outer segments with NAD+ and a purified NAD+:arginine ADP-ribosyltransferase from turkey erythrocytes resulted in approx. 60% inhibition of GTPase activity. Inhibition was dependent on both enzyme and NAD+, and was potentiated by the non-hydrolysable GTP analogues guanosine 5'-[beta gamma-imido]triphosphate (p[NH]ppG) and guanosine 5'-[beta gamma-methylene]triphosphate (p[CH2]ppG). The transferase ADP-ribosylated both the T alpha and T beta subunits of purified transducin. T alpha (39 kDa), after ADP-ribosylation, migrated as two distinct peptides with molecular masses of 42 kDa and 46 kDa on SDS/polyacrylamide-gel electrophoresis. T beta (36 kDa), after ADP-ribosylation, migrated as a 38 kDa peptide. With purified transducin subunits, it was observed that the GTPase activity of ADP-ribosylated T alpha, reconstituted with unmodified T beta gamma and photolysed rhodopsin, was decreased by 80%; conversely, reconstitution of T alpha with ADP-ribosyl-T beta gamma resulted in only a 19% inhibition of GTPase. Thus ADP-ribosylation of T alpha, the transducin subunit that contains the guanine nucleotide-binding site, has more dramatic effects on GTPase activity than does modification of the critical 'helper subunits' T beta gamma. To elucidate the mechanism of GTPase inhibition by transferase, we studied the effect of ADP-ribosylation on p[NH]pp[3H]G binding to transducin. It was shown previously that modification of transducin by choleragen, which like transferase ADP-ribosylates arginine residues, did not affect guanine nucleotide binding. ADP-ribosylation by the transferase, however, decreased p[NH]pp[3H]G binding, consistent with the hypothesis that choleragen and transferase inhibit GTPase by different mechanisms.  相似文献   

5.
Formation of reduced nicotinamide adenine dinucleotide peroxide   总被引:1,自引:0,他引:1  
Incubation of NADH at neutral and slightly alkaline pH leads to the gradual absorption of 1 mol of H+. This uptake of acid requires oxygen and mainly yields anomerized NAD+ (NAD+), with only minimal formation od acid-modified NADH. The overall stoichiometry of the reaction is: NADH + H+ + 1/2O2 leads to H2O + NAD+, with NADH peroxide (HO2-NADH+) serving as the intermediate that anomerizes and breaks down to give NAD+ and H2O2. The final reaction reaction mixture contains less than 0.1% of the generated H2O2, which is nonenzymically reduced by NADH. The latter reaction is inhibited by catalase, leading to a decrease in the overall rate of acid absorption, and stimulated by peroxidase, leading to an increase in the overall rate of acid absorption. Although oxygen can attack NADH at either N-1 or C-5 of the dihydropyridine ring, the attack appears to occur primarily at N-1. This assignment is based on the inability of the C-5 peroxide to anomerize, whereas the N-1 peroxide, being a quaternary pyridinium compound, can anomerize via reversible dissociation of H2O2. The peroxidase-catalyzed oxidation of NADH by H2O2 does not lead to anomerization, indicating that anomerization occurs prior to the release of H2O2. Chromatography of reaction mixtures on Dowex 1 formate shows the presence of two major and several minor neutral and cationic degradation products. One of the major products is nicotinamide, which possibly arises from breakdown of nicotinamide-1-peroxide. The other products have not been identified, but may be derived from other isomeric nicotinamide peroxides.  相似文献   

6.
Choleragen exerts its effect on cells through activation of adenylate cyclase. Choleragen initially interacts with cells through binding of the B subunit of the toxin to the ganglioside GM1 on the cell surface. Subsequent events are less clear. Patching or capping of toxin on the cell surface may be an obligatory step in choleragen action. Studies in cell-free systems have demonstrated that activation of adenylate cyclase by choleragen requires NAD. In addition to NAD, requirements have been observed for ATP, GTP, and calcium-dependent regulatory protein. GTP also is required for the expression of choleragen-activated adenylate cyclase. In preparations from turkey erythrocytes, choleragen appears to inhibit an isoproterenol-stimulated GTPase. It has been postulated that by decreasing the activity of a specific GTPase, choleragen would stabilize a GTP-adenylate cyclase complex and maintain the cyclase in an activated state. Although the holotoxin is most effective in intact cells, with the A subunit having 1/20th of its activity and the B subunit (choleragenoid) being inactive, in cell-free systems the A subunit, specifically the A1 fragment, is required for adenylate cyclase activation. The B protomer is inactive. Choleragen, the A subunit, or A1 fragment under suitable conditions hydrolyzes NAD to ADP-ribose and nicotinamide (NAD glycohydrolase activity) and catalyzes the transfer of the ADP-ribose moiety of NAD to the guandino group of arginine (ADP-ribosyltransferase activity). The NAD glycohydrolase activity is similar to that exhibited by other NAD-dependent bacterial toxins (diphtheria toxin, Pseudomonas exotoxin A), which act by catalyzing the ADP-ribosylation of a specific acceptor protein. If the ADP-ribosylation of arginine is a model for the reaction catalyzed by choleragen in vivo, then arginine is presumably an analog of the amino acid which is ADP-ribosylated in the acceptor protein. It is postulated that choleragen exerts its effects on cells through the NAD-dependent ADP-ribosylation of an arginine or similar amino acid in either the cyclase itself or a regulatory protein of the cyclase system.  相似文献   

7.
The steady-state kinetic mechanism of the reaction catalyzed by octopine dehydrogenase [N2-(1-carboxyethyl)-L-arginine:NAD+ oxidoreductase] was investigated at pH 6.9 and 9.2 by studies of substrate inhibition, analogue inhibition, and product inhibition. In the direction of octopine synthesis, the inhibition patterns in the presence of delta- guanidinovalerate and propionate show that NADH binds to the enzyme first followed by L-arginine and pyruvate which bind randomly. In the direction of octopine oxidation, the substrate patterns show that NAD binds to the enzyme before octopine in a rapid equilibrium fashion, and the product inhibition patterns show that the products L-arginine and pyruvate are released in a random fashion. Double, synergistic, substrate inhibition by L-arginine and pyruvate was shown to be due to binding (hypothetically of the imine) to the free enzyme and the enzyme-NAD complex. Furthermore, an alternate minor pathway was demonstrated which includes an enzyme-NADH-octopine complex and an enzyme-octopine complex.  相似文献   

8.
A putative long-chain mannitol-1-phosphate 5-dehydrogenase from Aspergillus fumigatus (AfM1PDH) was overexpressed in Escherichia coli to a level of about 50% of total intracellular protein. The purified recombinant protein was a approximately 40-kDa monomer in solution and displayed the predicted enzymatic function, catalyzing NAD(H)-dependent interconversion of d-mannitol 1-phosphate and d-fructose 6-phosphate with a specific reductase activity of 170 U/mg at pH 7.1 and 25 degrees C. NADP(H) showed a marginal activity. Hydrogen transfer from formate to d-fructose 6-phosphate, mediated by NAD(H) and catalyzed by a coupled enzyme system of purified Candida boidinii formate dehydrogenase and AfM1PDH, was used for the preparative synthesis of d-mannitol 1-phosphate or, by applying an analogous procedure using deuterio formate, the 5-[2H] derivative thereof. Following the precipitation of d-mannitol 1-phosphate as barium salt, pure product (>95% by HPLC and NMR) was obtained in isolated yields of about 90%, based on 200 mM of d-fructose 6-phosphate employed in the reaction. In situ proton NMR studies of enzymatic oxidation of d-5-[2H]-mannitol 1-phosphate demonstrated that AfM1PDH was stereospecific for transferring the deuterium to NAD+, producing (4S)-[2H]-NADH. Comparison of maximum initial rates for NAD+-dependent oxidation of protio and deuterio forms of D-mannitol 1-phosphate at pH 7.1 and 25 degrees C revealed a primary kinetic isotope effect of 2.9+/-0.2, suggesting that the hydride transfer was strongly rate-determining for the overall enzymatic reaction under these conditions.  相似文献   

9.
The kinetic mechanism of action of octopine dehydrogenase was investigated. This enzyme catalyses the reversible dehydrogenation of D-octopine to L-arginine and pyruvate, in the presence of nicotinamide-adenine dinucleotide. Initial velocity and product inhibition studies were carried out in both directions. Most of the results are consistent with a bi-ter sequential mechanism where NAD+ binds first to the enzyme followed by D-octopine, and the products are released in the order L-arginine, pyruvate and NADH. Various kinetic parameters were determined for each reactant at 33 degrees C, at pH 9.6 for NAD reduction, at pH 6.6 for NADH oxidation.  相似文献   

10.
Peroxisomal chain-shortening of prostaglandin F2 alpha   总被引:2,自引:0,他引:2  
We have recently reported that prostaglandin F2 alpha can be chain-shortened by isolated rat liver peroxisomes. In the present study it is further established by cell fractionation experiments that the enzymes involved in this reaction are localized to peroxisomes. Under the conditions employed, the highest activity was found in the light mitochondrial fraction. Further fractionation of the light mitochondrial fraction by sucrose density gradient centrifugation showed that the prostaglandin oxidation activity comigrated with peroxisomal marker enzymes. Di(2-ethylhexyl)phthalate treatment resulted in a tenfold increased capacity for the conversion of prostaglandin F2 alpha into tetranorprostaglandin F1 alpha. The reaction was not inhibited by KCN. The reaction was further characterized with respect to cofactor requirements. The prostaglandin oxidation was found to be completely dependent on NAD, CoA, ATP, Mg2+ and was stimulated by FAD. Incubation of prostaglandin E2 with peroxisomes resulted in conversion into several products. After alkaline hydrolysis, one of these was identified as tetranorprostaglandin B1.  相似文献   

11.
Present evidence suggests that skin is an important organ of prostaglandin metabolism. To clarify its role, the basic kinetics of 15-hydroxyprostaglandin dehydrogenase (PGDH) from rat skin were investigated with either NAD+ of NADP+ as co-substrate. Prostaglandin F2 alpha (PGF2 alpha) and prostaglandin E2 (PGE2) were used as substrates and preliminary studies were made of the inhibitory effects of the reduced co-substrates NADH and NADPH. A radiochemical assay was used in which [3H]PGF2 alpha or [14C]PGE2 were incubated with high-speed supernatant of rat skin homogenates. The substrate and products were then extracted by solvent partition, separated by t.l.c. and quantified by liquid-scintillation counting. At linear reaction rates and at an NAD+ concentration of 10 mM the mean apparent Km for PGF2 alpha was 24 microM with a mean apparent Vmax. of 9.8 nmol/s per litre of reaction mixture. For PGE2 the mean apparent Km was 8 microM, with a mean apparent Vmax, of 2.7 nmol/s per litre of reaction mixture. With NADP+ as a co-substrate at a concentration of 5 mM a mean apparent Km of 23 microM was obtained for PGF2 alpha with a mean apparent Vmax. of 5.2 nmol/s per litre. For PGE2 values of 7.5 microM and 3.0 nmol/s per litre were obtained respectively. These results show that skin contains NAD+- and NADP+-dependent PGDH. An important finding was that the NADP+-linked enzyme gave Km values for PGE2 that were considerably lower than those reported for NADP+-linked PGDH from other tissues. Furthermore, preliminary inhibition studies with the NAD+-linked PGDH system indicate that this enzyme is not only inhibited by NADH, but also by NADPH, a property not previously reported for NAD+-linked PGDH derived from other tissues.  相似文献   

12.
A metabolite with intracellular Ca2+-mobilizing activity can be produced by incubating NAD+ with extracts from sea urchin eggs. Structural determination indicates it is a cyclized ADP-ribose, and we have proposed cyclic ADP-ribose as a common name for it. In this study, we addressed the question of how widespread is the occurrence of the synthesizing enzyme for this NAD+ metabolite. Incubation of NAD+ with extracts prepared from rabbit liver resulted in a progressive increase in Ca2+ release activity which was monitored by a biological assay using sea urchin egg homogenates. The half-maximal concentration of NAD+ required was about 1 mM. The reaction was stereospecific, and the extracts were sensitive to protease treatment and heat, as well as alkaline pH of about 9.0, indicating the reaction was catalyzed by a protein. The active metabolite was purified by an identical high pressure liquid chromatography (HPLC) procedure used for cyclic ADP-ribose. Functionally, the liver metabolite behaved similarly to cyclic ADP-ribose. Both discharged the same Ca2+ stores in sea urchin egg homogenates with the same half-maximal effective concentrations. Both were active in inducing the cortical exocytosis reaction when microinjected into sea urchin eggs. That they are indeed identical compounds was demonstrated by structural analyses showing that they coeluted on a Partisil 5 SAX HPLC column and had identical 1H NMR spectra. Mass spectrometry indicated a mass of 540.0529 for the molecular ion (M - H)- of the liver metabolite, which was identical to within 0.74 ppm of cyclic ADP-ribose. Furthermore, their collisional activated decomposition mass spectra were virtually superimposable. Extracts from rabbit brain, heart, spleen, and kidney were all active in producing similar Ca2+-releasing metabolites which could be isolated by the same HPLC procedure and had similar elution times on both the mixed mode and the Partisil 5 SAX column. It is therefore apparent that the synthesizing enzyme for cyclic ADP-ribose is a very common enzyme.  相似文献   

13.
Amino acid racemases inherently catalyze the exchange of alpha-hydrogen of amino acids with deuterium during racemization in 2H2O. When the reactions catalyzed by alanine racemase (EC 5.1.1.1) and L-alanine dehydrogenase (EC 1.4.1.1), which is pro-R specific for the C-4 hydrogen transfer of NADH, are coupled in 2H2O, [4R-2H]NADH is exclusively produced. Similarly, [4S-2H]NADH is made in 2H2O with amino-acid racemase with low substrate specificity (EC 5.1.1.10) and L-leucine dehydrogenase (EC 1.4.1.9), which is pro-S specific. We have established a simple procedure for the in situ analysis of stereospecificity of C-4 hydrogen transfer of NADH by an NAD-dependent dehydrogenase by combination with either of the above two couples of enzymes in the same reaction mixture. When the C-4 hydrogen of NAD+ is fully retained after sufficient incubation, the stereospecificity of hydrogen transfer by a dehydrogenase is the same as that of alanine dehydrogenase or leucine dehydrogenase. However, when the C-4 hydrogen of NAD+ is exchanged with deuterium, the enzyme to be examined shows the different stereospecificity from alanine dehydrogenase or leucine dehydrogenase. Thus, we can readily determine the stereospecificity by 1H NMR measurement without isolation of the coenzymes and products.  相似文献   

14.
125I-labeled 1-(p-hydroxyphenyl) 2-guanidinoethane (N-guanyltyramine), previously used to assay for the bacterial toxin choleragen (Mekalanos, J.J., Collier, R.J. and Romig, W.R. (1979) J. Biol. Chem. 254, 5849-5854) was utilized to identify NAD:arginine ADPribosyltransferases in animal tissues. The use of this radiolabelled ADPribose acceptor, rather than radiolabelled NAD, would bypass the problem posed by the almost ubiquitous presence of enzymes that degrade NAD. With a homogeneous ADPribosyltransferase from turkey erythrocytes, NAD and 125I-labeled guanyltyramine as ADPribose acceptor, formation of ADPribosyl 125I-guanyltyramine was linear with time and enzyme concentration. The product was indistinguishable on both thin-layer and high-performance liquid chromatography from that formed by choleragen. Using 125I-guanyltyramine, ADPribosyltransferase activity was also demonstrated in crude turkey erythrocyte cytosolic and membrane fractions. When rat liver was fractionated, apparent activity was detected primarily in the microsomes. The NAD-dependent product of the microsomal reaction was, however, distinguished from the turkey erythrocyte transferase product by thin-layer and DEAE-Sephadex chromatography; this product had a retention time identical to that of free 125I on high-performance liquid chromatography. In addition to NAD, the microsomal deiodinase activity was supported by NADH, NADP and NADPH. Phenyl boronate selectively bound ADPribosyl 125I-guanyltyramine and other metabolites of 125I-guanyltyramine which were formed by microsomes in a NAD-dependent process. These metabolites were distinguished from ADPribosyl 125I-guanyltyramine by high-performance liquid chromatography. These results indicate that in some cases, for example, turkey erythrocyte cytosolic and membrane fractions, 125I-guanyltyramine can be used to quantify ADPribosyltransferases in crude mixtures, whereas in others, for example, rat liver microsomes, high-performance liquid chromatographic analysis must be used to identify products.  相似文献   

15.
15-Keto-13,14-dihydro 6-ketoprostaglandin E1 was positively identified by gas chromatography-mass spectrometry with negative-ion chemical ionisation detection from samples of rat kidney high-speed supernatant incubated with prostaglandin I2 in the presence of NAD+. A decreased formation of this product was observed when NAD+ was substituted with NADP+ and none was observed in the absence of nucleotide or substrate prostaglandin I2. Experiments with [9 beta-3H]prostaglandin I2 showed a time- and concentration-dependent loss of tritium which appeared as tritiated water, typical of reaction of [9 beta-3H]prostaglandin substrates with the enzyme, 9-hydroxyprostaglandin dehydrogenase. Time-course measurements of the appearance of tritiated water showed similar rates with 6-keto[9 beta-3H]prostaglandin F1 alpha and 15-keto-13,14-dihydro 6-keto[9 beta-3H]prostaglandin F1 alpha as substrates. These experiments suggest that the transformation of prostaglandin I2 and 6-ketoprostaglandin F1 alpha into the 15-keto-13,14-dihydro 6-ketoprostaglandin E1 catabolite occurs in this in vitro preparation via the corresponding 15-keto-13,14-dihydro catabolite of 6-ketoprostaglandin F1 alpha.  相似文献   

16.
尿苷二磷酸(uridine diphosphate,UDP)-葡萄糖醛酸是细胞内重要的糖基供体,参与多种代谢途径,也是体外进行糖基化反应的重要糖基供体,但其价格昂贵、工艺复杂,限制了其大量使用,无法满足生产需求。基于此,利用双酶偶联法氧化UDP-葡萄糖生成UDP-葡萄糖醛酸,并研究反应产物的合成情况。以UDP-葡萄糖为底物、烟酰胺腺嘌呤二核苷酸(nicotinamide adenine dinucleotide,NAD+)为辅因子,利用化脓性链球菌Streptococcus pyogenes源的尿苷二磷酸葡萄糖脱氢酶(UDP-glucose dehydrogenase,UGD)、猪源的乳酸脱氢酶(lactate dehydrogenase,LDH),双酶偶联催化合成UDP-葡萄糖醛酸,并通过高效液相色谱、质谱及核磁共振氢谱对反应产物进行检测,确定产物的结构及产物的生成量。结果表明,利用双酶偶联法氧化UDP-葡萄糖所得到的产物为UDP-葡萄糖醛酸。在UGD的作用下,氧化UDP-葡萄糖生成UDP-葡萄糖醛酸,同时辅因子NAD+在LDH的作用下实现循环再生,减少高能产物辅酶还原型烟酰胺腺嘌呤二核苷酸(reduced nicotinamide adenine dinucleotide,NADH)对反应的反馈抑制作用,产物的生成率约为60.17%。研究提高了产物UDP-葡萄糖醛酸产物生成量,为后续工业化制备提供了新思路。  相似文献   

17.
18.
Sea urchin eggs contain a small molecular weight heat-stable factor that confers cyanide-resistant NAD(P)H-O2 oxidoreductase activity on ovoperoxidase (Turner, E., Somers, C. E., and Shapiro, B. M. (1985) J. Biol. Chem. 260, 13163-13171), the enzyme responsible for cross-linking the extracellular protein coat (fertilization membrane) of the egg. Here we report the isolation of the active cofactor and its identification by ultraviolet, NMR, and mass spectroscopy as a new sulfur-containing amino acid derivative, 1-methyl-alpha N,alpha N-dimethyl-4-thiohistidine, or ovothiol. Ovothiol reacts slowly with atmospheric oxygen or rapidly with micromolar concentrations of H2O2 to form ovothiol disulfide, which is inactive as a cofactor for the ovoperoxidase NAD(P)H oxidase. Reduced active ovothiol is regenerated by treatment with disulfide reductants and shows significant differences in its ultraviolet and NMR spectra from oxidized ovothiol. The oxidoreductase activity of the ovoperoxidase/ovothiol system is similar to that previously characterized with crude cofactor preparations; it is greatly enhanced by Mn2+ and is relatively insensitive to CN-, compared to the peroxidase activity of ovoperoxidase. The ovothiol content of eggs is estimated as 1.8 pmol/egg or an intracellular concentration of 6.8 mM. This concentration exceeds the amount of reductant needed for the CN-(-)insensitive oxygen consumption following fertilization and used in the production of H2O2 for fertilization membrane cross-linking. Whether ovothiol is involved in the cross-linking reaction, protects the egg from damage from H2O2, or has another role in development remains unclear.  相似文献   

19.
Summary Choleragen exerts its effects on cells through the activation of adenylate cyclase. The initial event appears to be the binding of the B subunit of the toxin to ganglioside GM1 on the cell surface, following which there is a delay prior to activation of adenylate cyclase. Patching and capping of the toxin on the cell surface, perhaps involved in the internalization of the enzymatically active subunit, may be occuring during this time. The activation of adenylate cyclase, which is catalyzed by the A1 peptide of choleragen, does not require the B subunit or ganglioside GM1. The A1 peptide catalyzes the transfer of ADP-ribose from NAD to an amino acid, probably arginine, in a 42 000 dalton membrane protein. This protein appears to be the GTP-binding component (or G/F factor) of the adenylate cyclase system and is cruical to the regulation of cyclase activity by hormones such as epinephrine. ADP-ribosylation of the G/F factor is enhanced by GTP and, in some systems, by a cytosolic factor. GTP is also required for stabilization and optimal catalytic function of the choleragen-activated cyclase. Calmodulin, a calcium-binding protein, is necessary for expression of catalytic activity of the toxin-activated adenylate cyclase in brain and other tissues. The ADP-ribosyltransferase activity required for activation of the cyclase is an intrinsic property of the A1 peptide of choleragen which is expressed only after the peptide is released from the holotoxin by reduction of a single disulfide bond. In the absence of cellular components, choleragen catalyzes the ADP-ribosylation of small guanidino compounds such as arginine as well as peptides and proteins that contain arginine. It is assumed, therefore, that the site of ADP-ribosylation in the natural acceptor protein is an arginine or similar amino acid. When guanidino compounds are not present as ADP-ribose acceptors, choleragen hydrolyzes NAD to ADP-ribose and nicotinamide at a considerably slower rate. E. coli heat-labile enterotoxin (LT) is very similar to choleragen in structure and function. It consists of two types of subunits, A and B, with sizes comparable to those of the A and B subunits of choleragen. Binding of LT to the cell surface is enhanced by prior incorporation of GM1 but not other gangliosides; the oligosaccharide of GM1 specifically interacts with LT and its B subunit. The A subunit of LT exhibits ADP-ribosyltransferase activity following activation by thiol to release the A1 peptide. The A subunit of LT can be isolated in an ‘unnicked’ form and thus requires, in addition to reduction by a thiol, proteolytic cleavage to generate the active A1 peptide. Like choleragen, LT uses guanidino compounds as model ADP-ribose acceptors and catalyzes the ADP-ribosylation of a 42 000 dalton protein in cell membrane prepatations. ADP-ribosyltransferases that use arginine as ADP-ribose acceptors are not restricted to bacterial systems; such an enzyme has been purified to apparent homogeneity (>500 000-fold) from turkey erythrocytes. Based on a subunit molecular weight of 28 000, its turnover number with arginine as the ADP-ribose acceptor is considerably higher than that of either toxin. Although with low molecular weight guanidino derivatives the substrate specificity of the enzyme is similar to that of choleragen, with protein substrates it clearly differs. The physiological role of the turkey erythrocyte transferase remains to be established.  相似文献   

20.
Cytoplasmic Ca2+ signals are transferred to the mitochondria and activate the Krebs cycle. We have compared the efficiency of this process for two Ca2+ mobilising agonists, PGF2alpha and ATP (acting at metabotropic P2 receptors) in rat luteal cells. [Ca2+]c, [Ca2+]m and mitochondrial NAD(P)H were monitored by means of microspectrofluorimetry and confocal microscopy. While both agonists caused similar elevations of [Ca2+]c, changes in NAD(P)H were larger in response to PGF2alpha than to ATP. PGF2alpha more effectively increased NAD(P)H level also in mouse luteal cells. PGF2alpha caused a faster rate of rise of NAD(P)H fluorescence than ATP when reoxidation was prevented with rotenone, suggesting a faster rate of NAD(P)+ reduction. The NAD(P)H response to both agonists was dependent on the mobilisation of stored Ca2+. We found no difference in the efficacy of transmission of the [Ca2+]c signal to mitochondria in response to PGF2alpha and ATP. Raising [Ca2+]c with ionomycin increased the NAD(P)H signal, which was further raised by PGF2alpha but not by ATP. These data suggest that PGF2alpha potentiates the Ca2+-induced stimulation of mitochondrial metabolism by a Ca2+-independent mechanism and shows that agonists may modulate mitochondrial function differentially through a novel process beyond the simple transfer of Ca2+ from ER to mitochondria.  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