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.
Trehalose-6-phosphate (T-6-P) synthetase activity in extracts of Dictyostelium discoideum has been reexamined in an effort to resolve discrepancies between the results of previous studies (R. Roth and M. Sussman (1966). Biochim. Biophys. Acta, 122, 225; K. A. Killick and B. E. Wright (1972). J. Biol. Chem., 247, 2967). We find that T-6-P synthetase is not cold sensitive as reported by Killick and Wright (1972), is not present in bacterial-grown vegetative cells (though subject to some modulation by other nutritional conditions), and is not in our hands unmasked or activated by ammonium sulfate fractionation. We conclude that the pattern of T-6-P synthetase accumulation and disappearance during fruiting body construction in D. discoideum is as originally described by R. Roth and M. Sussman (1968). J. Biol. Chem., 243, 5081) and confirmed elsewhere (P. C. Newell et al. (1972). J. Mol. Biol., 63, 373; R. W. Brackenbury et al. (1974). J. Mol. Biol., 90, 529; B. D. Hames and J. M. Ashworth (1974). Biochem. J., 142, 301).  相似文献   

2.
The pH dependence of proton uptake upon binding of NADH to porcine heart mitochondrial malate dehydrogenase (l-malate: NAD+ oxidoreductase, EC 1.1.1.37) has been investigated. The enzyme has been shown to exhibit a pH-dependent uptake of protons upon binding NADH at pH values from 6.0 to 8.5. Enzyme in which one histidine residue has been modified per subunit by the reagent iodoacetamide (E. M. Gregory, M. S. Rohrbach, and J. H. Harrison, 1971, Biochim. Biophys. Acta253, 489–497) was used to establish that this specific histidine residue was responsible for the uptake of a proton upon binding of NADH to the native enzyme. It has also been established that while there is no enhancement of the nucleotide fluorescence upon addition of NADH to the iodoacetamide-modified enzyme, NADH is nevertheless binding to the modified enzyme with the same stoichiometry as with native enzyme. The data are discussed in relation to the involvement of the essential histidine residue in the catalytic mechanism of “histidine dehydrogenases” recently proposed by Lodola et al. (A. Lodola, D. M. Parker, R. Jeck, and J. J. Holbrook, 1978, Biochem. J.173, 597–605) and the catalytic mechanism of “malate dehydrogenases” recently proposed by L. H. Bernstein and J. Everse (1978, J. Biol. Chem.253, 8702–8707).  相似文献   

3.
In disagreement with reported observation by Suhara and her colleagues (K. Suhara, S. Takemori, M. Katagiri, K. Wada, H. Kobayashi, and H. Matsubara, 1975, Anal. Biochem.68, 632–636) we found that more than 90% of labile sulfur was liberated from adrenodoxin within 5 min at 22°C. This rate was faster than those of spinach and clostridial ferredoxins, a result also at variance with Suhara's observation. At low temperature, the reaction was clearly biphasic, and spinach ferredoxin showed a similar profile. In the absence of zinc acetate, activation energies of the decomposition reaction of iron-sulfur center of OH? were obtained as 39, 26, and 11 kcal/mol for adrenal, spinach, and clostridial ferredoxins, respectively. The adrenal reaction became faster as the dipole moment of the solvent increased. In the presence of 4 m urea and 1 m KCl, the rate was enhanced by approximately 26-fold, relative to the reaction without the addition of urea. In conclusion, the liberation reaction of adrenal labile sulfur with alkaline zinc reagent is fast at 22°C, indicating no need for modification of the original method (T. Kimura and K. Suzuki, 1967, J. Biol. Chem.242, 485–491; P. E. Brumly, R. W. Miller, and V. Massey, 1965, J. Biol. Chem.240, 2222–2228).  相似文献   

4.
Treatment of malic enzyme with arginine-specific reagents phenylglyoxal or 2,3-butanedione results in pseudo-first-order loss of oxidative decarboxylase activity. Inactivation by phenylglyoxal is completely prevented by saturating concentrations of NADP+, Mn2+, and substrate analog hydroxymalonate. Double log plots of pseudo-first-order rate constant versus concentration yield straight lines with identical slopes of unity for both reagents, suggesting that reaction of one molecule of reagent per active site is associated with activity loss. In parallel experiments, complete inactivation is accompanied by the incorporation of four [14C]phenylglyoxal molecules, and the loss of two arginyl residues per enzyme subunit, as determined by the colorimetric method of Yamasaki et al (R. B. Yamasaki, D. A. Shimer, and R. E. Feeney (1981) Anal. Biochem., 14, 220–226). These results confirm a 2:1 ratio for the reaction between phenylglyoxal and arginine (K. Takahashi (1968) J. Biol. Chem., 243, 6171–6179) and yield a stoichiometry of two arginine residues reacted per subunit for complete inactivation, of which one is essential for enzyme activity as determined by the statistical method of Tsou (C. L. Tsou (1962) Acta Biochim. Biophys. Sinica, 2, 203–211) and the Ray and Koshland analysis (W. J. Ray and D. E. Koshland (1961) J. Biol. Chem., 236, 1973–1979). Amino acid analysis of butanedione-modified enzyme also shows loss of arginyl residues, without significant decrease in other amino acids. Modification by phenylglyoxal does not significantly affect the affinity of this enzyme for NADPH. Binding of l-malate and its dicarboxylic acid analogs oxalate and tartronate is abolished upon modification, as is binding of the monocarboxylic acid α-hydroxybutyrate. The latter result indicates binding of the C-1 carboxyl group of the substrate to an arginyl residue on the enzyme.  相似文献   

5.
The recent assertion of J. Diguiseppi and I. Fridovich (1980, Arch. Biochem. Biophys., 203, 145–150) that Fe-EDTA does not catalyze superoxide dismutation is disputed. By directly observing superoxide generated during pulse radiolysis, we have confirmed the results of a previous study (G. J. McClune, J. A. Fee, G. A. McClusky, and J. T. Groves, 1977, J. Amer. Chem. Soc., 99, 5220–5222) which concluded that Fe-EDTA catalyzed superoxide dismutation. We also demonstrate that the reaction of Fe(II)-EDTA, formed during catalyzed superoxide dismutation, with cytochrome c, the probe molecule in the cytochrome c/xanthine oxidase/xanthine assay system for superoxide dismutase activity, is sufficiently rapid (H. L. Hodges, R. A. Holwerda, and H. B. Gray, 1974, J. Amer. Chem. Soc., 96, 3132–3137) to obscure the weak catalysis of superoxide dismutation by Fe-EDTA.  相似文献   

6.
Using nine different l-aminoacyl-4-nitroanilides and four different dipeptidyl-4-nitroanilides, aminopeptidases and dipeptidyl aminopeptidases active at pH 7.5 and (or) pH 5.5 in logarithmically growing and stationary-phase cells of Saccharomyces cerevisiae were searched for. Ion-exchange chromatography was used to separate the proteins of the soluble cell extract. Besides the three already-characterized aminopeptidases—aminopeptidase I (P. Matile, A. Wiemken, and W. Guyer (1971) Planta (Berlin)96, 43–53; J. Frey and K. H. Röhm (1978) Biochim. Biophys. Acta527, 31–41), aminopeptidase II (J. Frey and K. H. Röhm (1978) Biochim. Biophys. Acta527, 31–41; J. Knüver (1982) Thesis, Fachbereich Chemie, Marburg, FRG), and aminopeptidase Co (T. Achstetter, C. Ehmann, and D. H. Wolf (1982) Biochem. Biophys. Res. Commun.109, 341–347)—12 additional aminopeptidase activities are found in soluble cell extracts eluting from the ion-exchange column. These activities differ from the characterized aminopeptidases in one or more of the parameters such as charge, size, substrate specificity, inhibition pattern, pH optimum for activity and regulation. Also, a particulate aminopeptidase, called aminopeptidase P, is found in the nonsoluble fraction of disintegrated cells. Besides the described particulate X-prolyl-dipeptidyl aminopeptidase (M. P. Suarez Rendueles, J. Schwencke, N. Garcia-Alvarez and S. Gascon (1981) FEBS Lett.131, 296–300), three additional dipeptidyl aminopeptidase activities of different substrate specificities are found in the soluble extract.  相似文献   

7.
This note considers sampling theory for a selectively neutral locus where it is supposed that the data provide nucleotide sequences for the genes sampled. It thus anticipates that technical advances will soon provide data of this form in volume approaching that currently obtained from electrophoresis. The assumption made on the nature of the data will require us to use, in the terminology ofKimura (Theor. Pop. Biol.2, 174–208 (1971)), the “infinite sites” model of Karlin and McGregor (Proc. Fifth Berkeley Symp. Math. Statist. Prob.4, 415–438 (1967)) rather that the “infinite alleles” model of Kimura and Crow (Genetics49, 174–738 (1964)). We emphasize that these two models refer not to two different real-world circumstances, but rather to two different assumptions concerning our capacity to investigate the real world. We compare our results where appropriate with corresponding sampling theory of Ewens (Theor. Pop. Biol.3, 87–112 (1972)) for the “infinite alleles” model. Note finally that some of our results depend on an assumption of independence of behavior at individual sites; a parallel paper byWatterson (submitted for publication (1974)) assumes no recombination between sites. Real-world behavior will lie between these two assumptions, closer to the situation assumed by Watterson than in this note. Our analysis provides upper bounds for increased efficiency in using complete nucleotide sequences.  相似文献   

8.
Exact and approximate expressions are obtained for the probability that the most frequent allele is oldest, in neutral allele models in which all mutations produce new alleles. The higher the mutation rate, the less likely is it that the most frequent allele would be oldest. The results are in agreement with simulation studies by Ewens and Gillespie (1974) (Theor. Popul. Biol.6, 35–57), and limit the range of validity of a suggestion made by Crow (1972) (J. Hered.63, 306–316) with respect to the statistical testing of the neutral allele hypothesis.  相似文献   

9.
Eight neutral oligosaccharide fractions were obtained from the pooled urine of two patients with mannosidosis by Bio-Gel P2 and Bio-Gel P4 column chromatography. The structures of seventeen oligosaccharides were determined by monosaccharide composition analysis, methylation studies, acetolysis, Smith degradation, and 13C NMR analysis. Three of the proposed structures, Manα1-3Manβ1-4GlcNAc, Manα1-2Manα1-3Manβ1-4GlcNAc, and Manα1-2Manα1-2Manα1-3Manβ1-4GlcNAc are identical to those first published by Norden et al. (N. E. Norden, A. Lundblad, S. Svennson, P. A. Ockerman, and S. Autio, 1973. J. Biol. Chem.248, 6210–6215; N. E. Norden, A. Lundblad, S. Svennson, and S. Autio, 1974. Biochemistry13, 871–874). Thirteen of them, Manα1-3Manα1-6(Manα1-3)-Manβ1-4GlcNAc, Manα1-3Manα1-6(Manα1-2Manα1-3)Manβ1-4GlcNAc, and 11 isomers of (Manα1-2)0–4[Manα1-6(Manα1-3)Manα1-6(Manα1-3)Manβ1-4GlcNAc], are the same as those first published by Yamashita et al. (K. Yamashita, Y. Tachibana, K. Mihara, S. Okada, H. Yabuuchi, and A. Kobata, 1980, J. Biol. Chem.255, 5126–5133); a tetrasac-charide, Manα1-6(Manα1-3)Manβ1-4GlcNAc, is newly reported and several other structural possibilities are proposed.  相似文献   

10.
The reaction kinetics of acetyl-coenzyme A carboxylase purified from developing castor oil seeds have been examined. On the basis of the substrate interaction and product inhibition results, a hybrid ping-pong mechanism is proposed. This type of mechanism demands that the active site of the enzyme be separated into two functionally distinct catalytic sites. The carboxybiotin intermediate formed at one site by the hydrolysis of ATP swings to the second site where acetyl-CoA is carboxylated to form malonyl-CoA. This hybrid rapid-equilibrium random bi bi uni uni ping-pong mechanism which includes the formation of three abortive complexes, E · HCO3? · ADP, E · HCO3? · Pi and E · Pi · Pi, is analogous to the hybrid ping-pong mechanism previously described for methylmalonyl-CoA transcarboxylase (D. B. Northrop (1969) J. Biol. Chem., 244, 5808) and pyruvate carboxylase (R. E. Barden, C-H. Fung, M. F. Utter, and M. C. Scrutton (1972) J. Biol. Chem., 247, 1323).  相似文献   

11.
The adenylate cyclase system in the plasma membrane of fat cells contains regulatory components that either stimulate or inhibit activity in response to ligands acting at the cell surface. GTP is required for both the stimulation by various hormones (catecholamines and peptide hormones) and the inhibition by adenosine. We have analyzed the effects of high-energy radiation on the stimulatory and inhibitory processes and conclude that these processes are mediated by structures of different functional size. Moreover, the fat cell cyclase system, when analyzed under conditions in which the inhibitory action of GTP is minimally expressed, displays targets of the same size as those previously observed for those involved in the activation of the hepatic enzyme by glucagon and guanine nucleotides (W. Schlegel, E. S. Kempner, and M. Rodbell, 1979, J. Biol. Chem.254, 5168–5176). These findings extend our recent evidence for the nonidentity of the two GTP-mediated processes (D. M. F. Cooper, W. Schlegel, M. C. Lin, and M. Rodbell, 1979, J. Biol. Chem.254, 8927–8931).  相似文献   

12.
Murine plasmacytoma endoplasmic reticulum which has been freed of ribosomes by EDTA treatment is capable of the cotranslational proteolytic processing of representative λ12, and k immunoglobulin light chain precursors. Messenger RNA fractions from the MOPC-104E, MOPC-315, and MOPC-46B tumor lines were used to direct the synthesis of the light chain precursors in a cell-free system derived from Krebs II ascites cells. The precursor cleavage activity of the plasmacytoma membranes is comparable in activity and in characteristics to that of two well-defined membrane preparations: Krebs II ascites intracellular membranes (E. Szczesna and I. Boime, 1976, Proc. Nat. Acad. Sci. USA73, 1179–1183) and EDTA-treated rough endoplasmic reticulum from canine pancreas (34., 35., J. Cell Biol.67, 852–862). The efficiency of the cleavage reaction appears to be dependent upon the precursor being utilized as a substrate. An assay suitable for a preliminary characterization of the plasmacytoma membrane preparations is described.  相似文献   

13.
14.
Utilizing a temperature-sensitive mutant of Escherichia coli K-12 defective in the coupling of metabolic energy to active transport, we have demonstrated that the uptake systems for arabinose, galactose, valine, histidine, and glutamine, which are sensitive to the osmotic shock treatment of L. A. Heppel (1965) (J. Biol. Chem.240, 3685), are all totally defective at the nonpermissive temperature (42 °C) whereas the intracellular ATP levels increase twofold. Phosphate bond energy alone is therefore not sufficient to energize the transport of these substrates. We have confirmed the findings of E. A., Berger and L. A. Heppel (1974) (J. Biol. Chem. 249, 7747) regarding a severe arsenate I inhibition of the uptake of substrates belonging to osmotic shock-sensitive transport systems and therefore conclude that both ATP and a functional ecf gene product are required for the coupling of energy to the transport of these solutes.  相似文献   

15.
Migrating cells possess surface glycosyltransferase activity toward extracellular substrates, and the appearance of enzyme activity coincides with the onset of cellular migration (Shur, 1977a, Shur, 1977b, Develop. Biol.58, 23–39, 40–55; E. A. Turley and S. Roth, 1979, Cell17, 109–115). In this paper, surface glycosyltransferases were examined during normal and TT mutant mesenchyme migration. Of six glycosyltransferases that were assayed, only galactosyltransferase was present at significant levels on the cell surface, despite the presence of a variety of intracellular glycosyltransferases. All controls have been performed to show clearly the enzyme activity was cell surface localized. In both normal and TT embryos, surface galactosyltransferase activity was localized, by autoradiography, primarily to migrating mesenchymal cells, and to a lesser degree, to presumptive neural epithelium. During primitive streak formation, putative TT embryos were devoid of surface galactosyltransferase activity. However, as development progressed, the TT level of activity eventually exceeded wild-type levels by two- to sixfold and was evident in TT tissues prior to the onset of microscopic pathology. Other surface enzymes assayed did not show any TT-dependent increase in activity. The extracellular galactosyl acceptors were not chloroform:methanol soluble, and glycopeptides prepared by exhaustive Pronase digestion were excluded from Sephadex G-50. This large galactosylated glycoconjugate was readily digestable with endo-β-galactosidase, and, therefore, is similar to the poly-N-acetyllactosamine chains previously identified on early embryonic tissues (A. Kapadia, T. Feizi, and M. J. Evans, 1981, Exp. Cell. Res.131, 185–195; T. Muramatsu, G. Gachelin, M. Damonneville, C. Delarbre, and F. Jacob, 1979, Cell18, 183–191; A. Heifetz, W. J. Lennarz, B. Libbus, and Y. -C. Hsu, 1980, Develop. Biol.80, 398–408). These results support an involvement of surface galactosyltransferases in mesenchyme formation and during migration on poly-N-acetyllactosamine substrates.  相似文献   

16.
Sulfhydryl oxidase (glutathione-oxidizing activity) is closely associated with γ-glutamyltransferase (γ-glutamyl transpeptidase) in skim milk membranes. Similar close association of the two enzymatic activities in kidney membranes has led to the recent proposal that glutathione-oxidizing activity can be attributed to the action of γ-glutamyltransferase, itself, in generating cysteinylglycine which, in turn, catalyzes sulfhydryl group oxidation (O. W. Griffith and S. S. Tate, 1980, J. Biol. Chem.255, 5011–5014). However, a previously published procedure for the isolation of highly purified sulfhydryl oxidase from skim milk membranes (V. G. Janolino and H. E. Swaisgood, 1975, J. Biol. Chem.250, 2532–2538) leads to the effective separation of the two activities. Quantitative chromatographic analyses of GSH, GSSG, and Glu levels revealed that the highly purified sulfhydryl oxidase preparation catalyzes the direct oxidation of GSH to GSSG without detectable cleavage of the γ-glutamyl peptide bond. These results were confirmed by monitoring the time course of substrate disappearance and product formation using high-performance liquid chromatography. Conversely, a supernatant fraction enriched in γ-glutamyltransferase activity displayed no sulfhydryl group-oxidizing activity. 6-Diazo-5-oxo-l-norleucine selectively inhibited the transferase in crude preparations containing both sulfhydryl oxidase and γ-glutamyltransferase. It is concluded that sulfhydryl oxidase and γ-glutamyltransferase activities are distinct and separable.  相似文献   

17.
Mitochondrial sequences have been identified within a set of cloned complementary DNAs that had been copied from poly(A)+RNA of two embryonic stages of Xenopus laevis (Dworkin and Dawid, 1980, Dworkin and Dawid, 1980, Develop. Biol.76, 435–448 and 449–464). Mitochondrial sequences were found to be highly abundant in gastrula stage poly(A)+RNA sequences; in tadpole RNA their relative abundance is reduced severalfold. Mitochondrial sequences account for the most abundant poly(A)+RNA molecules in the gastrula population. The high abundance of mitochondrial RNA in early stages may be the consequence of the accumulation of large numbers of mitochondria in the egg.  相似文献   

18.
The dual wavelength assay technique (H. R. Levy, and G. H. Daouk, 1979, J. Biol. Chem.254, 4843–4847) is used to examine the rates of the NADP- and NAD-linked reactions of Leuconostoc mesenteroides glucose 6-phosphate dehydrogenase simultaneously under various conditions. Inhibition by ATP, MgATP2?, acetyl-CoA, and palmitoyl-CoA is greatly diminished at high glucose 6-P concentration which favors the NAD-linked reaction. Increasing NADPHNADP+ concentration ratios inhibit the NADP-linked, but stimulate the NAD-linked reaction. The selective effects of glucose 6-P and the NADPHNADP+ concentration ratio, which cannot be detected by conventional assays, are explained in terms of the differing kinetic mechanisms for the NADP-linked and NAD-linked reactions previously described (C. Olive, M. E. Geroch, and H. R. Levy, 1971, J. Biol. Chem.246, 2047–2057). It is proposed that these effects constitute the mechanism whereby the nucleotide specificity of the amphibolic glucose 6-phosphate dehydrogenase from Leuconostoc mesenteroides is regulated.  相似文献   

19.
This report presents electron microscopic evidence of statistically significant changes in the microtubule number concentration and length distribution after the attainment of monomer-polymer equilibrium (“steady state”). We also extend previous theoretical work on polymer redistribution (F. Oosawa, 1970, J. Theor. Biol.27, 69–86).  相似文献   

20.
The Lowry method (G. H. Lowry, N. J. Rosebrough, A. L. Farr, and R. J. Randall, 1951, J. Biol. Chem.193, 265–275) for protein concentration measurement has been automated to permit assay of samples with concentrations from 1 to 400 μg/ml. Calibration with solutions of bovine serum albumin resulted in a nonlinear (quadratic) curve. The quantity of color developed in the assay was found to be strongly dependent on the concentration of the Folin-Ciocalteu phenol reagent. Color yield peaked sharply at a reagent concentration 40% lower than that used in the Lowry procedure. Optimization of the reagent concentration is necessary to obtain maximum sensitivity from the Lowry assay.  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