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.
Two methods were used for the quantitation of S-adenosylmethionine decarboxylase protein. The first involved titrating the active site of the enzyme by reduction of the Schiff base between 3H-decarboxylated S-adenosylmethionine and the pyruvate prosthetic group with sodium cyanoborohydride. The second method was radioimmunoassay with rabbit antiserum which was used to determine the total immunoreactive enzyme protein. It was found that the increased S-adenosylmethionine decarboxylase activity produced in rat prostate by treatment with alpha-difluoromethylornithine and in both prostate and liver by methylglyoxal bis(guanylhydrazone) were due entirely to increases in the amount of enzyme protein. The ratio of enzyme activity to protein (measured by either method) remained constant in rats treated with the drugs. Treatment with 2% alpha-difluoromethylornithine in the drinking water for 3 days increased prostatic S-adenosylmethionine decarboxylase protein by 5-fold. A substantial part, but not all, of this increase could be accounted for by a slowing of the rate of degradation of the enzyme. The half-life for loss of activity and titratable protein after inhibition of protein synthesis by cycloheximide was increased from 35 to 108 min by treatment with alpha-difluoromethylornithine. However, the half-life for loss of immunoreactive protein which was considerably longer was only increased from 139 to 213 min. The molecular weight of the S-adenosylmethionine decarboxylase subunit determined by immunoblotting was 32,000, and no smaller immunoreactive fragments were detected. These results indicate that spermidine depletion produced by alpha-difluoromethylornithine affects the degradation of S-adenosylmethionine decarboxylase at an early step involving the loss of the active site without substantial breakdown of the protein.  相似文献   

2.
Acetylation of decarboxylated S-adenosylmethionine by mammalian cells   总被引:2,自引:0,他引:2  
Decarboxylated S-adenosylmethionine was found to be a substrate for the nuclear acetyltransferases that act on polyamines and on histones. The rate of acetylation of decarboxylated S-adenosylmethionine was more than twice that of spermidine at saturating substrate concentrations, and decarboxylated S-adenosylmethionine was an active inhibitor of the acetylation of histones by nuclear extracts from rat liver. The acetylation of decarboxylated S-adenosylmethionine occurred in vivo in SV-3T3 cells exposed to the ornithine decarboxylase inhibitor 2-(difluoromethyl)ornithine. The decline in putrescine and spermidine brought about by exposure to 2-(difluoromethyl)ornithine was found to be accompanied by a large rise in the content of both decarboxylated S-adenosylmethionine and acetylated decarboxylated S-adenosylmethionine. These results indicate that decarboxylated S-adenosylmethionine is metabolized not only in the well-known reactions in which it serves as an aminopropyl donor for polyamine biosynthesis but also by acetylation in reaction with acetyl coenzyme A. Furthermore, the inhibition of histone acetylation by decarboxylated S-adenosylmethionine could contribute to the biological effects brought about by inhibitors of ornithine decarboxylase.  相似文献   

3.
Effects of S-adenosyl-1,8-diamino-3-thiooctane on polyamine metabolism   总被引:3,自引:0,他引:3  
A E Pegg  K C Tang  J K Coward 《Biochemistry》1982,21(20):5082-5089
Exposure of mammalian cells (transformed mouse fibroblasts or rat hepatoma cells) to S-adenosyl-1,8-diamino-3-thiooctane produced profound changes in the intracellular polyamine content. Putrescine was increased and spermidine was decreased, consistent with the inhibition of spermidine synthase by this compound, which is a potent and specific "transition-state analogue inhibitor" of the isolated enzyme in vitro. The spermine content of the cells was increased by exposure to this drug presumably since spermine synthase was able to use a greater proportion of the available decarboxylated S-adenosylmethionine when spermidine synthase was inhibited. The decarboxylated S-adenosylmethionine content rose substantially because the activity of S-adenosylmethionine decarboxylase was increased in response to the decline in spermidine. These results indicate that S-adenosyl-1,8-diamino-3-thiooctane is taken up by mammalian cells and is an effective inhibitor of spermidine synthase in vivo and that S-adenosylmethionine decarboxylase is regulated by the content of spermidine, but not of spermine. The growth of SV-3T3 cells was substantially reduced in the presence of S-adenosyl-1,8-diamino-3-thiooctane at concentrations of 50 microM or greater. Such inhibition was reversed by the addition of spermidine but not by putrescine. When SV-3T3 cells were exposed to 5 mM alpha-(difluoromethyl)ornithine and 50 microM S-adenosyl-1,8-diamino-3-thiooctane, the content of all polyamines was reduced. Putrescine and spermidine declined by more than 90% and spermine by 80%. Such cells grew very slowly unless spermidine was added.  相似文献   

4.
Antibodies were produced in rabbits to homogeneous mouse kidney ornithine decarboxylase and used to determine the amount of this protein present in kidney extracts by a competitive radioimmunoassay procedure. The labeled ligand for this assay was prepared by reacting renal ornithine decarboxylase with [5-3H] alpha-difluoromethylornithine, an enzyme-activated irreversible inhibitor. The sensitivity of the assay was such that 1 ng of protein could be quantitated and the binding to ornithine decarboxylase of a macromolecular inhibitor (antizyme) or alpha-difluoromethylornithine did not affect the reaction. It was found that treatment of female mice with testosterone produced a 400-fold increase in ornithine decarboxylase protein in the kidney within 4-5 days. Exposure to cycloheximide or to 1,3-diaminopropane led to a rapid disappearance of the protein which paralleled the loss of enzyme activity. There was no sign of any immunoreactive but enzymatically inactive form of mouse kidney ornithine decarboxylase under any of the conditions investigated. The results indicate that fluctuations of the enzyme activity in this organ are mediated via changes in the amount of enzyme protein rather than by post-translational modifications or interaction with inhibitors or activators.  相似文献   

5.
Treatment of tobacco liquid suspension cultures with methylglyoxal bis(guanylhydrazone) (MGBG) an inhibitor of S-adenosylmethionine decarboxylase, resulted in a dramatic overproduction of a 35-kDa peptide on sodium dodecyl sulfate-polyacrylamide gel electrophoresis (Malmberg, R.L., and McIndoo, J. (1983) Nature 305, 623-625). MGBG treatment also resulted in a 20-fold increase in the activity of S-adenosylmethionine decarboxylase. Purification of S-adenosylmethionine decarboxylase from MGBG-treated cultures revealed that the overproduced 35-kDa peptide and S-adenosylmethionine decarboxylase are identical. Precursor incorporation experiments using [3H] methionine and [35S]methionine revealed that MGBG does not induce any increased synthesis of S-adenosylmethionine decarboxylase but rather stabilizes the protein to proteolytic degradation. The half-life of the enzyme activity was increased when MGBG was present in the growth medium. In addition to stabilizing S-adenosylmethionine decarboxylase, MGBG also resulted in the rapid and specific loss of arginine decarboxylase activity with little effect ornithine decarboxylase. The kinetics of this effect suggest that arginine decarboxylase synthesis was rapidly inhibited by MGBG. Exogenously added polyamines had little effect on ornithine decarboxylase, whereas S-adenosylmethionine and arginine decarboxylase activities rapidly diminished with added spermidine or spermine. Finally, inhibition of ornithine decarboxylase was lethal to the cultures, whereas inhibition of arginine decarboxylase was only lethal during initiation of growth in suspension culture.  相似文献   

6.
Bennett EM  Ekstrom JL  Pegg AE  Ealick SE 《Biochemistry》2002,41(49):14509-14517
S-Adenosylmethionine decarboxylase has been implicated in cell growth and differentiation and is synthesized as a proenzyme, which undergoes autocatalytic cleavage to generate an active site pyruvoyl group. In mammals, S-adenosylmethionine decarboxylase is active as a dimer in which each protomer contains one alpha subunit and one beta subunit. In many higher organisms, autocatalysis and decarboxylation are stimulated by putrescine, which binds in a buried site containing numerous negatively charged residues. In contrast, plant S-adenosylmethionine decarboxylases are fully active in the absence of putrescine, with rapid autocatalysis that is not stimulated by putrescine. We have determined the structure of the S-adenosylmethionine decarboxylase from potato, Solanum tuberosum, to 2.3 A resolution. Unlike the previously determined human enzyme structure, the potato enzyme is a monomer in the crystal structure. Ultracentrifugation studies show that the potato enzyme is also a monomer under physiological conditions, with a weak self-association constant of 6.5 x 10(4) M(-)(1) for the monomer-dimer association. Although the potato enzyme contains most of the buried charged residues that make up the putrescine binding site in the human enzyme, there is no evidence for a putrescine binding site in the potato enzyme. Instead, several amino acid substitutions, including Leu13/Arg18, Phe111/Arg114, Asp174/Val181, and Phe285/His294 (human/potato), provide side chains that mimic the role of putrescine in the human enzyme. In the potato enzyme, the positively charged residues form an extensive network of hydrogen bonds bridging a cluster of highly conserved negatively charged residues and the active site, including interactions with the catalytic residues Glu16 and His249. The results explain the constitutively high activity of plant S-adenosylmethionine decarboxylases in the absence of putrescine and are consistent with previously proposed models for how putrescine together with the buried, negatively charged site regulates enzyme activity.  相似文献   

7.
A number of nucleosides related to S-adenosylmethionine were tested for their inhibitory action on three enzymes involved in the biosynthesis of polyamines. The particular objective of the experiments was to determine whether any of the compounds could be used as selective inhibitors of the synthesis of spermine by spermine synthase. None of the nucleosides examined were potent inhibitors of S-adenosylmethionine decarboxylase. 5'-[(3-Aminopropyl)amino]-5'-deoxyadenosine dihydrochloride was quite a strong inhibitor of spermidine synthase (I50 of 7 microM) but was more than an order of magnitude less active than S-adenosyl-1,8-diamino-3-thiooctane, which is a mechanism-based inhibitor of this enzyme. 5'-[(3-Aminopropyl)amino]-5'-deoxyadenosine also inhibited spermine synthase with an I50 of 17 microM, but more selective inhibition of spermine synthase was produced by 9-[6(RS),8-diamino-5,6,7,8-tetradeoxy-beta-D-ribo-octofuranosyl]-9 H-purin-6- amine (I50 of 12 microM) and by dimethyl(5'-adenosyl)sulfonium perchlorate (I50 of 8 microM) since these compounds were much less active against spermidine synthase. Both 9-[6(RS),8-diamino-5,6,7,8-tetradeoxy-beta-D-ribo-octofuranosyl]-9 H-purin-6- amine and dimethyl(5'-adenosyl)sulfonium perchlorate were able to reduce the synthesis of spermine in SV-3T3 cells, but there was a compensatory increase in the concentration of spermidine, and there was no effect on cell growth. These results and those from experiments in which these spermine synthesis inhibitors were combined with inhibitors of spermidine synthase and ornithine decarboxylase indicated that the cells compensated for the inhibition of the aminopropyltransferases by increasing the production of decarboxylated S-adenosylmethionine and putrescine.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Ornithine decarboxylase has been purified 1,500-fold to homogeneity from a spe2 mutant of Saccharomyces cerevisiae which lacks S-adenosylmethionine decarboxylase and is derepressed for ornithine decarboxylase. The ornithine decarboxylase is a single polypeptide (Mr = 68,000) and requires a thiol and pyridoxal phosphate for activity. Addition of 10(-4) M spermidine and 10(-4) M spermine to the growth medium reduces the activity of the enzyme by 90% in 4 h. However, immunoprecipitation studies showed that the extracts of polyamine-treated cells contain as much enzyme protein as normal cell extracts. This loss of ornithine decarboxylase activity is probably due to a post-translational modification of enzyme protein because we found no evidence for any inhibitor of activity in the polyamine-treated cells.  相似文献   

9.
The contrasting control of lysosomal proteinases, protein turnover and proliferation was studied in 3T3 and SV-3T3 (SV-40-virus-transformed 3T3) cells. 1. In 3T3 cells, net protein accumulation proceeded from 5%/h (doubling time, T(d)=14h) in growing cells to 0%/h as cells became quiescent. SV-3T3 cells never ceased to gain protein, but rather decreased their protein accumulation rate from 6-7%/h (T(d)=10-12h) to 2%/h (T(d)=35-40h) just before culture death in unchanged medium. 2. In both cell types the rates of protein synthesis per unit of protein (a) were proportional to the initial serum concentration from 0 to 6%, and (b) declined under progressive depletion of undefined serum growth factors. In depleted growth medium, leucine incorporation per unit of protein in 3T3 and SV-3T3 cells declined to almost equal synthetic rates while the 3T3 cell existed in a steady state of zero net gain, and the SV-3T3 cell continued to gain protein at a rate of 2%/h. 3. Whereas a large fraction of the control of 3T3-cell net protein accumulation can be accounted for by an increase in degradation from 1%/h to 3%/h, the SV-3T3 cell did not exhibit a growth-related increase in degradation appreciably above 1%/h. 4. Thus, by using first-order kinetics, the continued net protein accumulation of the transformed cell can be accounted for by a failure to increase protein degradation, whereas fractional synthesis can be made to decline to a rate similar to that in the quiescent non-transformed cell. 5. Upon acute serum deprivation, both cell types similarly exhibited small rapid increases in proteolysis independent of cell growth state or lysosomal enzyme status. 6. The 3T3 cell increased its lysosomal proteinase activity in conjunction with increase in the growth-state-dependent proteolytic mechanism; however, the SV-3T3 cell failed to increase lysosomal proteinases or the growth-state-dependent proteolytic mechanism.  相似文献   

10.
1. The induction of ornithine decarboxylase activity in mouse 3T3 fibroblasts or an SV-40 transformed 3T3 cell line by serum was prevented by addition of the naturally occurring polyamines putrescine (butane-1,4-diamine) and spermidine. Much higher concentrations of these amines were required to fully suppress ornithine decarboxylase activity in the transformed SV-3T3 cells than in the 3T3 fibroblasts. 2. Synthetic alpha omega-diamines with 3--12 carbon atoms also prevented the increase in ornithine decarboxylase activity induced by serum in these cells. The longer chain diamines were somewhat more potent than propane-1,3-diamine in this effect, but the synthetic diamines were less active than putrescine in the 3T3 cells. There was little difference between the responses of 3T3 and SV-3T3 cells to the synthetic diamines propane-1,3-diamine and heptane-1,7-diamine. 3. These results are discussed in relation to the control of polyamine synthesis in mammalian cells.  相似文献   

11.
We have recently isolated, without using any inhibitors, a mutant of Chinese hamster ovary cell line which greatly overproduces ornithine decarboxylase in serum-free culture. Addition of polyamines (putrescine, spermidine, or spermine, 10 microM) or ornithine (1 mM), the precursor of polyamines, to the culture medium of these cells caused a rapid and extensive decay of ornithine decarboxylase activity. At the same time the activity of S-adenosylmethionine decarboxylase showed a less pronounced decrease. Notably, the polyamine concentrations used were optimal for growth of the cells and caused no perturbation of general protein synthesis. Spermidine and spermine appeared to be the principal regulatory amines for both enzymes, but also putrescine, if accumulated at high levels in the cells, was capable of suppressing ornithine decarboxylase activity. The amount of ornithine decarboxylase protein (as measured by radioimmunoassay) declined somewhat more slowly than the enzyme activity, but no more than 10% of the loss of activity could be ascribed to post-translational modifications or inhibitor interaction. Some evidence for inactivation through ornithine decarboxylase-antizyme complex formation was obtained. Gel electrophoretic determinations of the [35S]methionine-labeled ornithine decarboxylase revealed a rapid reduction in the synthesis and acceleration in the degradation of the enzyme after polyamine additions. No decrease in the amounts of the two ornithine decarboxylase-mRNA species, hybridizable to a specific cDNA, was detected, suggesting that polyamines depressed ornithine decarboxylase synthesis by selectively inhibiting translation of the message.  相似文献   

12.
Total poly(A)-containing mRNA was isolated from rat ventral prostate and translated in a reticulocyte lysate system. The proteins corresponding to S-adenosylmethionine decarboxylase were precipitated with a specific antiserum to this protein. Two proteins were found; one having an Mr of 32,000, which corresponded to the subunit of this enzyme, and a larger protein of Mr 37,000. Immunopurification of polysomes with the antiserum to S-adenosylmethionine decarboxylase followed by mRNA extraction yielded an mRNA preparation which was 10-30% pure mRNA for S-adenosylmethionine decarboxylase. The translation of this mRNA showed clearly that the protein of Mr 37,000 was a precursor of the Mr 32,000 S-adenosylmethionine decarboxylase subunit. Treatment with 2-difluoromethylornithine, which depletes cellular spermidine and is known to increase the content of S-adenosylmethionine decarboxylase protein, led to a 9-fold increase in the content of its mRNA, but treatment with methylglyoxal bis(guanylhydrazone) did not change the mRNA content.  相似文献   

13.
1. Polyamine concentrations were decreased in rats fed on a diet deficient in vitamin B-6. 2. Ornithine decarboxylase activity was decreased by vitamin B-6 deficiency when assayed in tissue extracts without addition of pyridoxal phosphate, but was greater than in control extracts when pyridoxal phosphate was present in saturating amounts. 3. In contrast, the activity of S-adenosylmethionine decarboxylase was not enhanced by pyridoxal phosphate addition even when dialysed extracts were prepared from tissues of young rats suckled by mothers fed on the vitamin B-6-deficient diet. 4. S-Adenosylmethionine decarboxylase activities were increased by administration of methylglyoxal bis(guanylhydrazone) (1,1'-[(methylethanediylidine)dinitrilo]diguanidine) to similar extents in both control and vitamin B-6-deficient animals. 5. The spectrum of highly purified liver S-adenosylmethionine decarboxylase did not indicate the presence of pyridoxal phosphate. After inactivation of the enzyme by reaction with NaB3H4, radioactivity was incorporated into the enzyme, but was not present as a reduced derivative of pyridoxal phosphate. 6. It is concluded that the decreased concentrations of polyamines in rats fed on a diet containing vitamin B-6 may be due to decreased activity or ornithine decarboxylase or may be caused by an unknown mechanism responding to growth retardation produced by the vitamin deficiency. In either case, measurements of S-adenosylmethionine decarboxylase and ornithine decarboxylase activity under optimum conditions in vitro do not correlate with the polyamine concentrations in vivo.  相似文献   

14.
Treatment of rats with large but sublethal doses of methyl glyoxal bis(guanylhydrazone), a potent in vitro inhibitor of animal S-adenosylmethionine decarboxylases, causes marked increases in the enzyme activity of extracts of kidney, ventral prostate, and testis which had been extensively dialyzed to remove any remaining drug. One day after administration of the inhibitor to female rats, the renal S-adenosylmethionine decarboxylase activity was 12 times the normal level and remained greatly enhanced for a further 24 hr. As indicated by decline in decarboxylase activity following depression of protein biosynthesis by injection of cycloheximide, the apparent half-life of the kidney enzyme in normal female rats is roughly 2 hr; in contrast, the apparent half-life of the enzyme is elevated to a value of more than 20 hr in animals that were previously treated with methyl glyoxal bis(guanylhydrazone). The increased renal S-adenosylmethionine decarboxylase activity following administration of the specific enzyme inhibitor in vivo may thus be due, at least in part, to stabilization of the enzyme against intracellular inactivation as a result either of direct combination of the enzyme protein with the inhibitor, or with substance(s) in the tissue whose levels are influenced by treatment with methyl glyoxal bis(guanylhydrazone).  相似文献   

15.
S-Adenosylmethionine decarboxylase belongs to a small class of amino acid decarboxylases that use a covalently bound pyruvate as a prosthetic group. It is an essential enzyme for polyamine biosynthesis and provides an important target for the design of anti-parasitic and cancer chemotherapeutic agents. We have determined the structures of S-adenosylmethionine decarboxylase complexed with the competitive inhibitors methylglyoxal bis(guanylhydrazone) and 4-amidinoindan-1-one-2'-amidinohydrazone as well as the irreversible inhibitors 5'-deoxy-5'-[N-methyl-N-[(2-aminooxy)ethyl]amino]adenosine, 5'-deoxy-5'-[N-methyl-N-(3-hydrazinopropyl)amino]adenosine, and the methyl ester analogue of S-adenosylmethionine. These structures elucidate residues important for substrate binding and show how those residues interact with both covalently and noncovalently bound inhibitors. S-Adenosylmethionine decarboxylase has a four-layer alphabeta betaalpha sandwich fold with residues from both beta-sheets contributing to substrate and inhibitor binding. The side chains of conserved residues Phe7, Phe223, and Glu247 and the backbone carbonyl of Leu65 play important roles in binding and positioning the ligands. The catalytically important residues Cys82, Ser229, and His243 are positioned near the methionyl group of the substrate. One molecule of putrescine per monomer is observed between the two beta-sheets but far away from the active site. The activating effects of putrescine may be due to conformational changes in the enzyme, to electrostatic effects, or both. The adenosyl moiety of the bound ligand is observed in the unusual syn conformation. The five structures reported here provide a framework for interpretation of S-adenosylmethionine decarboxylase inhibition data and suggest strategies for the development of more potent and more specific inhibitors of S-adenosylmethionine decarboxylase.  相似文献   

16.
Ornithine decarboxylase (L-ornithine carboxy-lyase, EC 4.1.1.17) has been purified from 3T3- and SV40-transformed 3T3 mouse fibroblasts by affinity chromatography, and the physicochemical properties of the two enzymes compared. Measured properties include molecular weight of the active species, subunit molecular weight and specific activity of the purified enzymes, kinetic parameters, thermostability, degradation rate in vivo and immunological cross-reactivity. Although crude extracts of the transformant possess more ornithine decarboxylase activity per mg of protein than the parent strain, there is no evidence for the appearance of an altered form of the enzyme in these cells. The results reported in the present paper indicate that the increased ornithine decarboxylase activity in the transformed cells is the result of higher enzyme biosynthesis de novo.  相似文献   

17.
1. The production of (14)CO(2) from S-adenosyl[carboxyl-(14)C]methionine by rat liver extracts was investigated. It was found that, in addition to the well-known cytosolic putrescine-activated S-adenosylmethionine decarboxylase, an activity carrying out the production of (14)CO(2) could be extracted from a latent, particulate or membrane-bound form by treatment with buffer containing 1% (v/v) Triton X-100 [confirming the report of Sturman (1976) Biochim. Biophys. Acta428, 56-69]. 2. The formation of (14)CO(2) by such detergent-solubilized extracts differed from that by cytosolic S-adenosylmethionine decarboxylase in a number of ways. The reaction by the solubilized extracts did not require putrescine and was not directly proportional to time of incubation or the amount of protein added. Instead, activity a showed a distinct lag period and was much greater when high concentrations of the extracts were used. The cytosolic S-adenosylmethionine decarboxylase was activated by putrescine, showed strict proportionality to protein added and the reaction proceeded at a constant rate. Cytosolic activity was not inhibited by homoserine or by S-adenosylhomocysteine, whereas the Triton-solubilized activity was strongly inhibited. 3. By using an acetone precipitate of Triton-treated homogenates as a source of the activity, it was found that decarboxylated S-adenosylmethionine was not present among the products of the reaction, although 5'-methylthioadenosine and 5-methylthioribose were found. Such extracts were able to produce (14)CO(2) when incubated with [U-(14)C]-homoserine, and (14)CO(2) production was greater when S-adenosyl[carboxyl-(14)C]methionine that had been degraded by heating at pH6 at 100 degrees C for 30min (a procedure known to produce mainly 5'-methylthioadenosine and homoserine lactone) was used as a substrate than when S-adenosyl[carboxyl-(14)C]methionine was used. 4. These results indicate that the Triton-solubilized activity is not a real S-adenosylmethionine decarboxylase, but that (14)CO(2) is produced via a series of reactions involving degradation of the S-adenosyl-[carboxyl-(14)C]methionine. It is probable that this degradation can occur via several pathways. Our results would suggest that part of the reaction occurs via the production of S-adenosylhomocysteine, which can then be converted into 2-oxobutyrate via the transsulphuration pathway, and that part occurs via the production of homoserine by an enzyme converting S-adenosylmethionine into 5'-methylthioadenosine and homoserine lactone.  相似文献   

18.
l-Tyrosine decarboxylase (EC 4.1.1.25) activity was induced in cell suspension cultures of Thalictrum rugosum Ait. and Eschscholtzia californica Cham. with a yeast polysaccharide preparation (elicitor). The highest l-tyrosine decarboxylase activity in extracts from 7-day-old cell cultures of E. californica was observed 5 hours after addition of 30 to 40 micrograms elicitor per gram cell fresh weight. The enzyme extracted from cells of E. californica was purified 1540-fold to a specific activity of 2.6 micromoles CO(2) produced per minute per milligram protein at pH 8.4 and 30 degrees C. Purified enzyme from T. rugosum showed a specific activity of 0.18 micromoles per minute per milligram protein. The purification procedure involved ammonium sulfate fractionation, anion-exchange fast protein liquid chromatography, ultrafiltration, and hydrophobic interaction chromatography. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed that the enzyme from the two plant cell cultures had subunits of identical molecular weight (56,300 +/- 300 daltons.  相似文献   

19.
3T3 cells have a large, pericellular coat which contains 30 times more hyaluronate than the amount of cell surface hyaluronate associated with simian virus 40-transformed 3T3 (SV-3T3) cells. On the other hand, SV-3T3 cells have high affinity binding sites for exogenously added hyaluronate, whereas 3T3 cells have much lower affinity sites. Removal of cell surface hyaluronate from SV-3T3 cells by treatment with hyaluronidase caused a reproducible increase in their maximum binding capacity for exogenous hyaluronate but no significant change in binding affinity or specificity. For 3T3 cells, however, the maximum amount of binding decreased and the affinity of binding increased after hyaluronidase treatment. When endogenous cell surface hyaluronate was labeled metabolically and then the cells incubated in the presence of exogenous unlabeled hyaluronate, the labeled cell surface hyaluronate was quantitatively displaced from the SV-3T3 cells but was not displaced from the 3T3 cells. Chondroitin sulfate and heparin did not displace cell surface hyaluronate from either cell type. Membranes isolated from SV-3T3 cells bound hyaluronate specifically and with high affinity, whereas membranes from 3T3 cells did not consistently bind a significant amount of hyaluronate. We conclude from these studies that the retention of endogenous hyaluronate on the surface of SV-3T3 cells is mediated by binding sites similar to those detected by the addition of exogenous hyaluronate, and the mechanism of retention of endogenous hyaluronate on the surface of 3T3 cells differs from SV-3T3 cells.  相似文献   

20.
E Gilboa  C L Prives  H Aviv 《Biochemistry》1975,14(19):4215-4220
SV-40 DNA sheared form was coupled in a stable covalent bond to cyanogen bromide activated Sepharose. Under the conditions used at least 80% of the DNA was bound to Sepharose. The T 1/2 of hybridization of 0.5 mug/ml of SV-40 cRNA to SV-40 DNA-Sepharose was 1 hr. This rate of hybridization is sufficiently rapid to purify SV-40 sequences from solutions containing as little as 0.05-0.1 mug/ml. Nonspecific hybridization of RNA is in the range of 0.1-0.2% of the total input RNA. The DNA-Sepharose is fairly stable and can be reused several times to purify RNA. The SV-40 DNA-Sepharose was used to select large quantities of virus specific RNA from SV-40 infected BS-C-1 cells. The virus specific RNA when added to cell-free extracts from wheat germ was shown to direct the synthesis of the major viral structural protein VP-1.  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