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 312 毫秒
1.
We previously have described a substance present in crude sonicates of L-929 cells which replaced ascorbate in vitro as a reductant for prolyl hydroxylase (B. Peterkofsky, D. Kalwinksy and R. Assad, 1980, Arch. Biochem. Biophys.199, 362–373). In the present study we found that almost 90% of the substance was particulate after differential centrifugation of stationary phase L-929 cell homogenates. The substance was not localized in nuclei or mitochondria and was found in the same fractions as microsomes, but these fractions also contained lysosomes and cell membranes. The reductant could not be solubilized from particles by Brij-35, indicating that it is an intrinsic component of a membrane rather than intracisternally located. The intramembranous cofactor, in the absence of ascorbate, participated in the in vitro hydroxylation of [4-3H]proline in radio-actively labeled, intracisternal unhydroxylated procollagen in isolated microsomes which also contained prolyl hydroxylase. Hydroxylation was determined by measuring tritiated water formed from release of the 4-trans tritium atom. Since it is unlikely that such participation could occur if the cofactor were located within the membrane of another subcellular organelle, we have concluded that it is in the same particle as prolyl hydroxylase and unhydroxylated procollagen, that is, the microsome. With the endogenous reductant the reaction was slower than with saturating ascorbate and was increased by NADH. Maximum hydroxylation with the endogenous reductant was close to that which could be achieved with ascorbate. These results provide strong evidence that the endogenous reductant alone can account for the phenomenon of ascorbate-independent proline hydroxylation in L-929 cells. As in the case of ascorbate, the microsomal reductant functioned only in the presence of α-ketoglutarate and Fe2+ and served as reductant for lysyl hydroxylase. It also was detected in the particulate fraction of virally transformed BALB 3T3 cells and in purified microsomes from bones of intact chick embryos. Since ascorbate could be taken up and concentrated in bone microsomes, it is unlikely that the endogenous reductant serves as an intermediary between ascorbate and intracisternal prolyl hydroxylase.  相似文献   

2.
An improved procedure was used to assay prolyl hydroxylase activity in both early-log and late-log L-929 fibroblasts grown on plastic surfaces. When 40 μg/ml of ascorbate was added to early-log phase cultures, the rate of hydroxy-[14C] proline synthesis increased 2-fold within 4 h, but there was no change in prolyl hydroxylase activity per cell. The results indicated therefore that ascorbate did not “activate” prolyl hydroxylase in the sense of converting inactive enzyme protein to active enzyme protein. Instead ascorbate appeared to increase hydroxyproline synthesis in early-log L-929 fibroblasts because the prolyl hydroxylase reaction in such cells was limited by the availability of ascorbate or a similar cofactor. When 40 μg/ml of ascorbate was added to late-log phase cultures, there was essentially no effect on the rate of hydroxyl[14C]-proline synthesis or prolyl hydroxylase activity. The late-log phase cells, however, contained three times more enzyme activity and about two times more immuno-reactive enzyme protein than early-log phase cells. In addition, the rate of protein synthesis per cell in late-log phase cells was only one-tenth the rate in early-log phase cells. The results suggested that as the cells grew to confluency, collagen polypeptides were more completely hydroxylated in part because the rate of polypeptide synthesis decreased and at the same time prolyl hydroxylase activity per cell increased. The results appear to provide an alternate explanation for previous observations on the effects of ascorbate and “crowding” on hydroxy[roline synthesis in cultures of L-929 fibroblasts.  相似文献   

3.
The incorporation of DL-3,4-dehydro[14C]proline into collagen and total protein of 3T3 cells occurred at approximately one-fifth the rate observed for L-[14C]proline. Addition of L-3,4-dehydroproline to the culture medium inhibited markedly the incorporation of [14C]glycine and L-[3H]lysine into the collagen of 3T3 cells, but there was only slight inhibition of the incorporation of the radiolabeled amino acids into total cellular proteins, indicating that the action of L-3,4-dehydroproline is specific for collagen. When 1 mM L-3,4-dehydroproline was added to the culture medium the [14C]hydroxyproline content was reduced 40% in the cell layer and 70% in the medium. The D isomer of 3,4-dehydroproline did not inhibit [14C]hydroxyproline formation. These findings indicate that L-3,4-dehydroline reduced the hydroxylation of the susceptible prolyl residues in the collagen molecule and the secretion of collagen from the cell. The reduction in the hydroxyproline content is probably related in part to a reduction in the activity of prolyl hydroxylase; when various mammalian cell cultures were exposed to 0.2 mM L-3,4-dehydroproline, the specific activity of prolyl hydroxylase was reduced markedly, while that of lysyl hydroproline, the specific activity of prolyl hydroxylase was reduced markedly, while that of lysyl hydroxylase was not affected. Under these conditions, cell growth and lactic dehydrogenase required protein synthesis. Removal of L-3,4-dehydroproline from the growth medium resulted in a time-dependent increase in the specific activity of prolyl hydroxylase.  相似文献   

4.
Reductant used as cofactor for the prolyl hydroxylase reaction, was measured by a tritium release assay modified from an enzyme assay by making all components of the assay system saturating except for the reductant, but including prolyl hydroxylase. Reduced glutathione (6 mm), which had little activity as a cofactor, and thymol (0.1 mm), an antioxidant which exhibited no cofactor activity at all, were required for optimal proline hydroxylation dependent on reducing cofactor, with thymol fulfilling the previously described requirement for catalase. Ascorbate, cysteine and 6,7-dimethyltetrahydropterin were active as cofactors, in descending order of activity at equimolar concentrations, and activity was concentration dependent for all of these compounds. Sonicates of stationary phase L-929 cells which exhibit ascorbate-independent proline hydroxylation in culture contained reducing cofactor which could replace ascorbate in the cofactor assay, while sonicates of log phase cells which exhibit an ascorbate requirement in culture contained about one-third or less of that amount. NADH and NADPH, which themselves have little or no activity as cofactor, increased the cofactor activity of log phase cell sonicates but had relatively little effect on the activity of stationary cell sonicates suggesting that the cofactor is in a more reduced state in stationary phase. Within 24 h after replating dense, stationary phase cell cultures at low density, conditions where cells return to ascorbate dependence, prolyl hydroxylase activity had decreased to one-fifth the original activity while the concentration of functional reducing cofactor had decreased to less than 1% of its original concentration, largely as a result of oxidation. Ascorbate was not present in L-929 cells sonicates and the levels of tetrahydropterin and cysteine in sonicates could not account for the amount of cofactor activity exhibited by the sonicates in the assay system. Treatment of L-929 cultures with aminopterin did not decrease ascorbate independence, suggesting that tetrahydrofolate did not contribute significantly to cellular proline hydroxylation. These results suggest that an unidentified reductant present in L-929 cells can account for ascorbate-independent proline hydroxylation and also regulate prolyl hydroxylase activity in these cells and that cellular levels of reduced pyridine nucleotides may regulate the reduction state of this substance.  相似文献   

5.
Inhibitors of purified, soluble prolyl hydroxylase (K. Majamaa et al. (1984) Eur. J. Biochem. 138, 239-245; K. Majamaa et al. (1986) J. Biol. Chem. 261, 7819-7823) were tested against isolated chick embryo bone microsomes containing intracisternal prolyl hydroxylase and its radiolabeled, unhydroxylated procollagen substrate. Two groups of inhibitors were used which consisted of pyridine-2-carboxylate and 1,2-dihydroxybenzene (catechol) derivatives. The 2,4- and 2,5-pyridine dicarboxylic acids, which are potent inhibitors of the soluble enzyme (Ki values 2 and 0.8 microM, respectively), were effective in the same concentration range against intracisternal prolyl hydroxylase, although their relative affinities were reversed. Inhibition by pyridine-2,4-dicarboxylate in the microsomal system was reversed by increasing the concentration of 2-oxoglutarate. Pyridine-2,4-dicarboxylic acid did not inhibit the uptake of 2-[14C]oxoglutarate into microsomes, so it appears likely that the inhibitor must traverse the microsomal membrane and act directly at the enzyme level. Pyridine-2-carboxylic acid was ineffective in the microsomal system at 1 mM whereas it is a relatively potent inhibitor of the soluble enzyme with a Ki of 25 microM. This finding suggests that the second carboxyl group of the pyridine carboxylate derivatives may be required for their transport into the microsomal lumen. In the soluble system, 3,4-dihydroxybenzoic acid and 1,2-dihydroxybenzene had been found to be competitive inhibitors with relatively low Ki values of 5 and 25 microM, respectively. In the microsomal system, half-maximal inhibition was obtained at approximately 50-100 microM and inhibition was not reversed by increasing the concentrations of either 2-oxoglutarate or ascorbate, alone or together. These results imply that in situ these compounds do not inhibit prolyl hydroxylase directly. Thus, the microsomal system can assess the accessibility of the intracisternal enzyme to potential inhibitors and offers an insight into the in cellulo potential of such compounds.  相似文献   

6.
Ascorbate was linked to protein folding a long time ago. At the first level of this connection, it had been shown that ascorbate functions as an essential cofactor in the hydroxylation enzymes involved in collagen synthesis. Although the hydroxylation reactions catalyzed by the members of the prolyl 4-hydroxylase family are considered to be ascorbate dependent, the hydroxylation of proline alone does not need ascorbate. Prolyl 4-hydroxylases participate in two catalytic reactions: one in which proline residues are hydroxylated, while 2-oxoglutarate is decarboxylated and molecular oxygen is consumed. This reaction is ascorbate independent. However, in another reaction, prolyl 4-hydroxylases catalyze the decarboxylation of 2-oxoglutarate uncoupled from proline hydroxylation but still needing molecular oxygen. At this time, ferrous iron is oxidized and the protein is rendered catalytically inactive until reduced by ascorbate. At the second level of the connection, the oxidation and the oxidized form of ascorbate, dehydroascorbate, is involved in the formation of disulfide bonds of secretory proteins. The significance of the dehydroascorbate reductase activity of protein disulfide isomerase was debated because protein disulfide isomerase as a dehydroascorbate reductase was found to be too slow to be the major route for the reduction of dehydroascorbate (and formation of disulfides) in the endoplasmic reticulum lumen. However, very recently, low tissue ascorbate levels and a noncanonical scurvy were observed in endoplasmic reticulum thiol oxidase- and peroxiredoxin 4-compromised mice. This novel observation implies that ascorbate may be involved in oxidative protein folding and creates a link between the disulfide bond formation (oxidative protein folding) and hydroxylation.  相似文献   

7.
T Kimura  D J Prockop 《Biochemistry》1982,21(22):5482-5488
[14C]Proline-labeled protocollagen, the unhydroxylated form of procollagen, was isolated from cartilage cells incubated with alpha, alpha'-dipyridyl. For examination of the initial steps in the hydroxylation of the protein, it was incubated in vitro with prolyl hydroxylase so that an average of 1.3-2.7 prolyl residues per chain was hydroxylated. The partially hydroxylated alpha chain were cleaved with cyanogen bromide, and the fragments were separated by polyacrylamide gel electrophoresis or column chromatography. The cyanogen bromide fragments were hydroxylated to the same degree. The results indicated, therefore, that in the initial hydroxylation of alpha chains in vitro, there was no preferential hydroxylation of any specific regions of the protein. In a second series of experiments, cartilage cells were incubated with [14C]proline and alpha, alpha'-dipyridyl so that prolyl hydroxylase in the cells was extensively, but not completely, inhibited. Partially hydroxylated alpha chains were isolated, and cyanogen bromide fragments of the alpha chains from the cells were assayed for hydroxy[14C]proline. The alpha chains contained an average of two residues of hydroxyproline per chain, and the cyanogen bromide fragments were hydroxylated to about the same degree. The results indicated, therefore, that when prolyl hydroxylase activity in cells is low relative to the rate at which pro alpha chains are synthesized, hydroxylation of prolyl residues occurs as it does in vitro, and there is no preferential hydroxylation of a specific region of the protein.  相似文献   

8.
New assays for the tyrosine hydroxylase and dopa oxidase activities of tyrosinase (EC 1.14.18.1) have been developed. The tyrosine hydroxylase assay uses L-[carboxy-14C]tyrosine as the substrate, 14CO2 is released from the products of the hydroxylation and further metabolism of L-[carboxy-14C]tyrosine by incubation with ferricyanide, and measured radiometrically. D-Dopa is a preferable cofactor to L-dopa for the assay. Dopa oxidase activity is measured spectrophotometrically. Dopaquinone, produced on the oxidation of L-dopa, reacts with Besthorn's hydrazone (3-methyl-2-benzothiazolinone hydrazone) to form a pink pigment with an absorbance maximum at 505 nm. Details of the optimisation of conditions for the assays and their specificities for the two enzyme activities are described.  相似文献   

9.
L-929 and 3T6 cells were conditioned to grow in a chemically defined medium lacking serum and ascorbate. Serum, when added, had a small stimulatory effect on the growth rate of the cells, but ascorbate had no effect either on the growth rate or on the rate of protein synthesis. These cells were also shown to lack gulonolactone oxidase activity and therefore could not synthesize their own ascorbate. Nevertheless, in the absence of serum and ascorbate both cell types were able to hydroxylate peptidyl proline to an appreciable extent. This suggests that reductants other than ascorbate can at least partially satisfy the requirement for a reductant in the prolyl hydroxylase reaction in vivo.  相似文献   

10.
L-929 and 3T6 cells were conditioned to grow in a chemically defined medium lacking serum and ascorbate. Serum, when added, had a small stimulatory effect on the growth rate of the cells, but ascorbate had no effect either on the growth rate or on the rate of protein synthesis. These cells were also shown to lack gulonolactone oxidase activity and therefore could not synthesize their own ascorbate. Nevertheless, in the absence of serum and ascorbate both cell types were able to hydroxylate peptidyl proline to an appreciable extent. This suggest that reductant other than ascorbate can at least partially satisfy the requirement for a reductant in the prolyl hydroxylase reaction in vivo.  相似文献   

11.
Collagen secretion by chick embryo fibroblasts was measured by incorporating [14C]proline into proteins and then analyzing the amount of collagen in the cell and medium separately by using purified bacterial collagenase. In order to produce varying levels of hydroxylation, cells were incubated with varying concentrations of ascorbate or with varying concentrations of α,α′-dipyridyl in the presence of saturating ascorbate. Ascorbate stimulated both the hydroxylation of proline in collagen and the secretion of collagen; the concentration of ascorbate required for half-maximal stimulation of both proesses was approximately 4.5 × 10?7, m. Since the cells could concentrate ascorbate 10-fold, this KM for proline hydroxylation is 100-fold lower than values reported for purified prolyl hydroxylase (Abbot, M. T., and Udenfriend, S. (1974) in Molecular Mechanisms of Oxygen Activation (Hayaishi, O., ed.), p. 173, Academic Press New York; Kivirikko K. I., et al. (1968) Biochim. Biophys. Acta, 151, 558–567). Conversely, α,ga′-dipyridyl inhibited both proline hydroxylation and collagen secretion; half-maximal inhibition of both processes was observed at 7 × 10?5, m. The results of the two types of experiments show that the secretion of collagen becomes directly proportional to proline hydroxylation when approximately 30% of the proline residues in collagen have been hydroxylated compared to maximal hydroxylation of 50%. Since the stability of triple-helical collagen at 37 °C has been shown to be dependent on the hydroxyproline content of the molecule (Rosenbloom, J., et al. (1973) Arch. Biochem. Biophys., 158, 478–484), we suggest that the observed proportionality between secretion and hydroxylation is a reflection of the increased amount of stable triple helical collagen at 37 °C. When the cells were incubated with a concentration of ascorbate that was saturating for secretion and hydroxylation, there was no significant activation of prolyl hydroxylase as measured in a cell-free extract. These experiments suggest that ascorbate effects collagen secretion by acting at the site of proline hydroxylation but not by increasing the activity of prolyl hydroxylase.  相似文献   

12.
There are two forms of prolyl hydroxylase in L-929 flbroblasts. One is the enzymatically active tetramer having two α- and two β-subunits. The other is monomeric cross-reacting protein which is enzymatically inactive but is structurally related to β-subunit of the enzyme. Cultured L-929 fibroblasts at mid-log phase were labeled by 3H-labeled amino acid mixture and the radioactivity was chased for 24 h while cells were harvested and plated at higher cell densities in cultures. The results indicated that both α-subunit of the tetrameric prolyl hydroxylase and cross-reacting protein were labeled, but the β-subunit of the tetrameric active prolyl hydroxylase was not labeled until the cells were crowded for 24 h. Using immunofluorescent techniques with antibodies directed against pure tetrameric prolyl hydroxylase, capping or patching was observed when the cells were incubated at 37 °C. Also, it was found that phagosomes prepared from L-929 flbroblasts contained about 30% of total enzyme protein as determined immunologically but contained no significant prolyl hydroxylase activities. Labeling cells with 125I by lactoperoxidase, cross-reacting protein was labeled but both α- and β-subunits of tetrameric active prolyl hydroxylase were not labeled. The results indicate that cross-reacting protein can be utilized as the precursor of β-subunit by the cells to form tetrameric active prolyl hydroxylase and that cross-reacting protein is found associated with cytoplasmic membranes.  相似文献   

13.
An improved procedure was developed to extract prolyl hydroxylase from tendon cells of chick embryos with detergent, and improved assays were developed for both the activity of the enzyme and the amount of enzyme protein. Freshly isolated tendon cells were found to contain approx. 100 mug of enzyme protein per 10(8) cells and 40-50% of the enzyme protein was active. When the cells were cultured, they were found to contain the same amount of enzyme protein but only 15-20% of the enzyme protein was active. Gel filtration of cell extracts indicated that the active form of prolyl hydroxylase in freshly isolated tendon cells and incultured tendon cells had the same apparent size and the same activity per mug of immunoreactive protein as enzyme which was shown to be a tetramer. The inactive form was found to have about the same apparent size as subunits of the enzyme. When freshly isolated cells were incubated for 2 h in the presence of 40 mug per ml of ascorbate, there was a slight increase in the rate of hydroxyproline synthesis. In cultured cells, ascorbate at a concentration of 40 mug per ml caused a 2-fold increase in the rate of hydroxyproline synthesis within 30 min. However, ascorbate did not icrease the activity of prolyl hydroxylase in extracts from either cell system. Therefore it appears that the influence of ascorbate on synthesis of procollagen hydroxyproline by the cells studied here must be ascribed to a cofactor effect on the hydroxylation reaction similar to that observed with purified enzyme, and it does not involve "activation" of inactive enzyme protein to active enzyme as has been observed in cultures of L-929 and 3T6 mouse fibroblasts.  相似文献   

14.
The control of ascorbic acid synthesis and turnover in pea seedlings   总被引:10,自引:0,他引:10  
The rate of ascorbate synthesis and turnover in pea seedling embryonic axes was investigated in relation to its pool size. Ascorbate accumulated in embryonic axes of germinating pea seeds which has been supplied with ascorbate. Incorporation of [U-14C]glucose into ascorbate after a 2 h labelling period was reduced by ascorbate loading for 3 h and 20 h, providing evidence that ascorbate biosynthesis is inhibited by endogenous ascorbate. Ascorbate turnover was estimated by following the metabolism of [1-14C]ascorbate over 2 h after ascorbate loading and by the rate of decrease of the ascorbate pool size after ascorbate loading. Ascorbate turnover rate, determined by [1-14C]ascorbate metabolism, increased as a linear function of pool size. The absolute turnover rate was higher in ascorbate-loaded embryonic axes but was always about 13% of the pool per hour. The initial rate of ascorbate turnover, estimated from the net decrease in pool size after ascorbate loading, also showed a similar turnover rate to that estimated from [1-14C]ascorbate metabolism. Ascorbate loading had no effect on ascorbate peroxidase, monodehydroascorbate reductase, dehydroascorbate reductase or glutathione reductase activity. Ascorbate oxidase activity decreased after ascorbate loading.  相似文献   

15.
The turnover time of prolyl hydroxylase in the skin of one-day-old rats treated in vivo with L-[14C]leucine is 22--34 hrs. Theoretically this time interval comprises two processes: i. e. incorporation of [14C]-Leu into the inactive enzyme precursor and transformation of the inactive from into the active enzyme.  相似文献   

16.
Prolyl hydroxylase activity in cultured L-929 cells was found to increase when cells grew from log phase to stationary phase and when cells were harvested at the mid-log phase and replated at higher cell densities. Cycloheximide and actinomycin D inhibited the cell density-dependent increase in prolyl hydroxylase activity indicating that the increase in prolyl hydroxylase activity required de novo synthesis of protein and RNA. Prolyl hydroxylase was purified from cultured L-929 cells and antibodies against the protein were raised in rabbits. The antibodies were used to demonstrate that L-929 cells contained two forms of prolyl hydroxylase: an enzymatically active, tetrameric form consisting of two alpha and two beta polypeptide chains and an enzymatically inactive form containing immunologically cross-reacting protein. The polypeptide chains alpha, beta and cross-reacting protein were obtained by immunoadsorption. Peptide map analysis indicated that cross-reacting protein was similar if not identical to beta in primary structure, and alpha was different from both beta and cross-reacting protein. The results suggested that the prolyl hydroxylase levels in cells or tissues may be regulated by new protein and/or RNA synthesis.  相似文献   

17.
Summary The action of hydralazine on collagen prolyl hydroxylation was studied in a cell culture system using WI-38 fibroblasts. The prolyl hydroxylation level was determined by a method involving the digestion of collagen by bacterial collagenase and the examination of specific peptides. The presence of low concentrations of hydralazine (0.2 mM) in both “young” and “old” fibroblast cultures strongly inhibited collagen prolyl hydroxylation. The degree of inhibition was greater in serum-deficient cultures. No significant improvement in the degree of hydroxylation was observed by increasing either ascorbate or iron levels in the hydralazine-containing cultures in which hydroxylation was inhibited. Some of the reported side effects of hydralazine seen in patients might be related to its inhibitory effects on mixed function oxidative (MFO) hydroxylation systems. While the ascorbate dependence of the prolyl hydroxylase system of WI-38 decreased with the “age” of the culture, hydralazine inhibition of hydroxylation was dramatic with cultures of all “ages”. This work was supported by NIH grants nos. AM15671, AM1675 and HD07376, and fellowship no. HD01998.  相似文献   

18.
An improved procedure was developed to extract prolyl hydroxylase from tendon cells of chick embryos with detergent, and improved assays were developed for both the activity of the enzyme and the amount of enzyme protein. Freshly isolated tendon cells were found to contain approx. 100 μg of enzyme protein per 108 cells and 40–50% of the enzyme protein was active. When the cells were cultured, they were found to contain the same amount of enzyme protein by only 15–20% of the enzyme protein was active. Gel filtration of cell extracts indicated that the active form of prolyl hydroxylase in freshly isolated tendon cells and in cultured tendon cells had the same apparent size and the same activity per μg of immunoreactive protein as enzyme which was shown to be a tetramer. The inactive form was found to have about the same apparent size as subunits of the enzyme.When freshly isolated cells were incubated for 2 h in the presence of 40 μg per ml of ascorbate, there was a slight increase in the rate of hydroxyproline synthesis. In cultured cells, ascorbate at a concentration of 40 μg per ml caused a 2-fold increase in the rate of hydroxyproline synthesis within 30 min. However, ascorbate did not increase the activity of prolyl hydroxylase in extracts from either cell system. Therefore it appears that the influence of ascorbate on synthesis of procollagen hydroxyproline by the cells studied here must be ascribed to a cofactor effect on the hydroxylation reaction similar to that observed with purified enzyme, and it does not involve “activation” of inactive enzyme protein to active enzyme as has been observed in cultures of L-929 and 3T6 mouse fibroblasts.  相似文献   

19.
The localization of collagen hydroxylysine galactosyl- and galactosyl-hydroxylysine glucosyltransferases in purified chick embryo bone microsomes was studied by differential solubilization with nonionic detergents. Brij-35 (polyoxyethylene 25-lauryl ether) which selectively releases intracisternal proteins, and Triton X-100, whose specificity varies with its concentration, were used in the presence or absence of high ionic strength NaCl. These methods were used previously to characterize prolyl hydroxylase as intracisternal and lysyl hydroxylase as mainly intramembranous. The distribution of both glycosyltransferases within microsomes was similar to that of lysyl hydroxylase; approximately 70-80% of their activities are intramembranous with the remainder intracisternal. Collagen hydroxylysine glucosyltransferase differed from prolyl and lysyl hydroxylase and the galactosyltransferase in that its activity in vitro was apparently inhibited by membrane vesicles, even in the presence of detergents at concentrations which permeabilize the membrane. Accurate measurement of its activity could be achieved only by its separation from vesicles after detergent treatment. The common location of the major portion of lysyl hydroxylase and the glycosyltransferase activities suggests that they may act as a multienzyme complex to preferentially modify certain lysyl residues in nascent procollagen chains as they traverse the membrane of the endoplasmic reticulum. Since these enzymes do not act on helical collagen, their physical separation from prolyl hydroxylase may ensure that modifications of lysine residues occur prior to formation of hydroxyproline, which stabilizes the helical form.  相似文献   

20.
The anthracyclines doxorubicin and daunorubicin were found to act as irreversible inhibitors of prolyl 4-hydroxylase. The reaction rate for enzyme from both chick and human origin was first order, the concentration of inhibitor giving 50% inhibition being 60 microM for both compounds after 1 h. The effect was dependent on the presence of iron ions in the reaction mixture. Inactivation could be prevented by addition of high concentrations of ascorbate, but not 2-oxoglutarate, before the inactivation period. The same results were obtained with competitive analogues of these cosubstrates. Lysyl hydroxylase from chick embryos was also susceptible to inactivation. Its activity was decreased by 50% after incubation for 1 h with a 150 microM concentration of the inhibitors. When chick-embryo prolyl 4-hydroxylase was incubated with [14-14C]doxorubicin, both enzyme subunits were radioactively labelled, about 70% of the total radioactivity being found in the alpha-subunit. Since the anthracyclines are known to undergo a redox reaction generating semiquinone radicals with Fe3+ only, the results suggest that the enzyme-bound iron ion is oxidized to a tervalent intermediate in uncoupled reaction cycles. The data also suggest that both enzyme subunits contribute to the catalytic site of prolyl 4-hydroxylase.  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