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.
The adenine nucleotides, 5'-AMP and 3',5'-cyclic AMP block L cells in the S-phase of the cell cycle. The intracellular level of cyclic AMP is reduced after incubation of cells with 5'-AMP, and rates of uridine transport are increased after incubation with either 5'-AMP or cyclic AMP. On the contrary, cyclic AMP levels are increased and uridine transport decreased in cells treated with an inhibitor of the cyclic AMP phosphodiesterase. This inhibitor partially reverses the growth-inhibitory effect of cyclic AMP, indicating that a breakdown product is the effective inhibitor of growth. The inhibition of cell growth induced by the adenine nucleotides is prevented by uridine, suggesting that the block in S is due to a lack of availability of pyrimidines.  相似文献   

2.
Summary We have measured by rapid kinetic techniques the zero-trans influx of hypoxanthine in various cell lines and its sensitivity to inhibition by uridine, dipyridamole, nitrobenzylthioinosine and nitrobenzylthiopurine. The results and those reported earlier divided the cells into two distinct groups. In mouse P388, L1210 and L929 cells uridine and hypoxanthine had little effect on the transport of each other, supporting the view that nucleosides and hypoxanthine are transported by different carriers. In these cells, hypoxanthine transport was also uniquely resistant to inhibition by dipyridamole (IC50 (50% inhibition dose) >30M). In Novikoff and HTC rat hepatoma, Chinese hamster ovary and Ehrlich ascites tumor cells, on the other hand, hypoxanthine and uridine inhibited the transport of each other about 50% at a concentration corresponding to the Michaelis-Menten constant of their transport, and hypoxanthine transport was strongly inhibited by dipyridamole (IC50=100 to 400nM). Although these results are compatible with the view that nucleosides and hypoxanthine are transported by a common carrier in these cells, this conclusion is not supported by the finding that uridine transport is strongly inhibited in some of these cell lines, as in first group of cells, by nitrobenzylthioinsine, whereas hypoxanthine transport is highly resistant in all cell lines tested. In contrast, the transport of both substrates is highly resistant to inhibition by nitrobenzylthiopurine. The Michaelis-Menten constants for uridine transport are about the same in all cell lines. The Michaelis-Menten constants for hypoxanthine transport are similar to those for uridine transport in some cell lines, but are much higher in others. This difference is unrelated to the sensitivity of uridine and hypoxanthine transport to inhibition by each other or dipyridamole.  相似文献   

3.
The adenine nucleotides of human red cells were labeled by incubation of the cells with [3H]adenosine. Then, the cells were incubated in Tris-saline with various supplements that cause the loss of cellular ATP, and the degradation products were quantitated as a function of time of incubation at 37 degrees C. Incubation of the cells with 2.5 or 5 mM iodoacetate, iodoacetamide or 1 mM HCHO in combination with 5 mM KF and 50 mM deoxyglucose, 50 mM D-glucose or 10 mM inosine was most efficient in depleting the cells of ATP (100% in 0.5-1 h) without causing cell lysis. In iodoacetate- and iodoacetamide-treated cells practically all catabolism of ATP occurred via ADP----AMP----IMP----inosine----hypoxanthine with hypoxanthine accumulating in the medium. In HCHO-treated cells and in cells incubated in Tris-saline or in Tris-saline with deoxyglucose with and without KF, a substantial proportion of ATP (up to 50%) was catabolized via ADP----AMP----adenosine----inosine----hypoxanthine. Under all conditions, AMP deamination and IMP and AMP hydrolysis were rate-limiting reactions. IMP degradation was more rapid in iodoacetamide- and HCHO-treated than in iodoacetate-treated red cells. It was also more rapid in fresh than in outdated red cells, and it was inhibited by Pi. Treatment with iodoacetamide and HCHO under ATP-depletion conditions resulted in a 60-80% inhibition of uridine transport by the cells. Treatment with iodoacetate or deoxyglucose plus KF had only minor effects on nucleoside transport; thus, cells treated in this manner might be useful for studying the transport of adenosine and deoxyadenosine under conditions were their phosphorylation is prevented.  相似文献   

4.
Exposure of 3T3-L1 adipocytes to 100 ng/ml of cholera toxin or 1 mM dibutyryl cyclic AMP caused a marked stimulation of deoxyglucose transport. A maximal increase of 10- to 15-fold was observed after 12-24 h of exposure, while 100 nM insulin elicited an increase of similar magnitude within 30 min. A short term exposure (4 h) of cells to cholera toxin or dibutyryl cyclic AMP resulted in a 3- to 4-fold increase in deoxyglucose transport which was associated with significant redistribution of both the HepG2/erythrocyte (GLUT1) and muscle/adipocyte (GLUT4) glucose transporters from low density microsomes to the plasma membrane fraction. Total cellular amounts of both transporter proteins remained constant. In contrast, cells exposed to cholera toxin or dibutyryl cyclic AMP for 12 h exhibited elevations in total cellular contents of GLUT1 (but not GLUT4) protein to about 1.5- and 2.5-fold above controls, respectively. Although such treatments of cells with cholera toxin (12 h) versus insulin (30 min) caused similar 10-fold enhancements of deoxyglucose transport, a striking discrepancy was observed with respect to the content of glucose transporter proteins in the plasma membrane fraction. While insulin elicited a 2.6-fold increase in the levels of GLUT4 protein in the plasma membrane fraction, cholera toxin increased the amount of this transporter by only 30%. Insulin or cholera toxin increased the levels of GLUT1 protein in the plasma membrane fraction equally (1.6-fold). Thus, a greater number of glucose transporters in the plasma membrane fraction is associated with transport stimulation by insulin compared to cholera toxin. We conclude that: 1) at early times (4 h) after the addition of cholera toxin or dibutyryl cyclic AMP to 3T3-L1 adipocytes, redistribution of glucose transporters to the plasma membrane appears to contribute to elevated deoxyglucose uptake rates, and 2) the stimulation of hexose uptake after prolonged treatment (12-18 h) of cells with cholera toxin may involve an additional increase in the intrinsic activity of one or both glucose transporter isoforms.  相似文献   

5.
The zero-trans uptake of uniformly and base-labeled inosine and uridine was measured a 25 degrees C in suspensions of Novikoff rat hepatoma cells, Chinese hamster ovary cells, mouse L cells, mouse S49 lymphoma cells and a purine-nucleoside phosphorylase-deficient subline thereof (NSU-1), and in monolayer culture of mouse 3T3 and L cells. The initial velocities of uptake of both nucleosides were about the same in all cell lines investigated, regardless of the position of the label or of the substrate concentration between 3 and 300 microM or whether or not the cells possessed uridine or purine-nucleoside phosphorylase activity. The kinetic parameters for the facilitated transport of uridine and inosine were also similar in phosphorylase positive and negative cell lines (K = 120--260 microM and V = 6--40 pmol/microliters cell water per s) and the transport activities of the cells exceeded their total phosphorylase activities by at least 10-fold for uridine and 1--2-fold for inosine. Chromatographic fractionation of the intracellular contents and of the culture fluid showed that the free nucleosides appeared intracellularly prior to and more rapidly than their phosphorolysis products. During the initial 20--60 s of uptake of U-14C-labeled nucleosides the rates of intracellular appearance of ribose-1-P and base were about the same. After several minutes of incubation, on the other hand, the main intracellular component was ribose-1-P whereas the base attained a low intracellular steady-state concentration and accumulated in the medium due to exit transport. Other nucleosides, dipyridamole and nitrobenzylthioinosine, specifically inhibited the transport of uridine and inosine, and depressed the intracellular accumulation of ribose-1-P and the formation of base commensurate with that inhibition. The data indicate that the metabolism of inosine and uridine by the various cell lines can be entirely accounted for by the facilitated transport of unmodified nucleoside into the cell followed by intracellular phosphorolysis.  相似文献   

6.
Nucleoside and nucleobase transport and metabolism were measured in ATP-depleted and normal Aedes albopictus mosquito cells (line C-7-10) by rapid kinetic techniques. The cells possess a facilitated diffusion system for nucleosides, which in its broad substrate specificity and kinetic properties resembles that present in many types of mammalian cells. The Michaelis-Menten constant for uridine transport at 28 degrees C is about 180 microM. However, the nucleoside transporter of the mosquito cells is resistant to inhibition by nmolar concentrations of nitrobenzylthioinosine and the cells lack high affinity nitrobenzylthioinosine binding sites. The cells also possess an adenine transporter, which is distinct from the nucleoside transporter. They lack, however, a hypoxanthine transport system and are deficient in hypoxanthine phosphoribosyltransferase activity, which explains their failure to efficiently salvage hypoxanthine from the medium. The cells possess uridine and thymidine phosphorylase activities and, in contrast to cultured mammalian cells, efficiently convert uracil to nucleotides. An adenosine-resistant variant (CAE-3-6) of the C-7-10 cell line is devoid of significant nucleoside transport activity but transports adenine normally. Residual entry of various nucleosides into these cells and of hypoxanthine and cytosine into wild type and mutant cells is strictly non-mediated. The rate of permeation of various nucleosides and of hypoxanthine into the CAE-3-6 cells is related to their hydrophobicity. Uridine permeation into CAE-3-6 cells exhibits an activation energy of about 20 kcal/mol. At high uridine concentrations permeation is sufficiently rapid to partly overcome the limitation in nucleoside salvage imposed by the nucleoside transport defect in these cells.  相似文献   

7.
Populations of Novikoff rat hepatoma cells (subline N1S1-67) were monitored for the rates of transport of various substrates and for their incorporation into acid-insoluble material as a function of the age of cultures of randomly growing cells in suspension as well as during traverse of the cells through the cell cycle. Populations of cells were synchronized by a double hydroxyurea block or by successive treatment with hydroxyurea and Colcemid. Kinetic analyses showed that changes in transport rates related to the age of cultures or the cell cycle stage reflecte alterations in the V max of the transport processes, whereas the Km remained constant, indicating that changes in transport rates reflect alterations in the number of functional transport sites. The transport sites for uridine and 2-deoxy-D-glucose increased continuously during traverse of the cells through the cell cycle, whereas those for choline and hypoxanthine were formed early in the cell cycle. Increases in thymidine transport sites were confined to the S phase. Synchronized cells deprived of serum failed to exhibit normal increases in transport sites, although the cells divided normally at the end of the cell cycle. Arrest of the cells in mitosis by treatment with Colcemid prevented any further increases in transport rates. The formation of functional transport sites was also dependent on de novo synthesis of RNA and protein. Inhibition of DNA synthesis in early S phase inhibited the increase in thymidine transport rates which normally occurs during the S phase, but had no effect on the formation of the other transport systems. Transport rates also fluctuated markedly with the age of the cultures of randomly growing cells, reaching maximum levels in the mid-exponential phase of growth. The transport systems for thymidine and uridine were rapidly lost upon inhibition of protein and RNA synthesis, and thus seem to be metabolically unstable, whereas the transport systems for choline and 2-deoxy-D-glucose were stable under the same conditions.  相似文献   

8.
Cytochalasin B competitively inhibits the transport of uridine and thymidine by Novikoff rat hepatoma cells growing in suspension culture with apparent Ki''s of 2 and 6 µM, respectively, but has no effect on the intracellular phosphorylation of the nucleosides. Choline transport is not affected by cytochalasin B. Results from pulse-chase experiments indicate that cytochalasin B has no direct effect on the synthesis of RNA, DNA, or uridine diphosphate-sugars. The inhibition of uridine and thymidine incorporation into nucleic acids by cytochalasin B is solely the consequence of the inhibition of nucleoside transport.  相似文献   

9.
Cytochalasin B competitively inhibits the transport of 2-deoxy-D-glucose and thymidine in a number of different cell lines (Novikoff rat hepatoma cells, mouse L, S180 and Ki-MSV-transformed BALB/3T3 cells, and human HeLa cells). The apparent Km values for the transport of these substrates as well as the apparent Ki values for the inhibition by cytochalasin B are very similar for the various cell lines, and the effect is readily and completely reversed by removal of the chemical. Thymidine transport by Chinese hamster ovary cells however, is little affected by cytochalasin B, whereas the transport of 2-deoxy-D-glucose, uridine and guanine by these cells is competitively inhibited to about the same extent as in other cell lines. In addition and concomitant with the inhibition of cytokinesis and an alteration in cell shape, cytochalasin B also impairs and delays the formation of functional transport sites for thymidine, guanine and choline in synchronized populations of Novikoff cells without affecting the apparent affinities of the transport systems for their substrates. This effect is unrelated to the direct inhibition of the transport processes, since the drug does not directly inhibit choline transport and has no effect on the formation of 2-deoxy-D-glucose transport sites in spite of the fact that it strongly inhibits the transport of this substrate. The inhibition of functional transport sites may be due to the induction of a structural alteration in the membrane by cytochalasin B which impairs the insertion of new proteins of certain but not all transport systems into the membrane.  相似文献   

10.
The initial rates of deoxy-D-glucose transport by cultures of growing and density-inhibited mouse embryo cells and lines of mouse cells transformed spontaneously or after infection by murine leukemia virus or murine sarcoma virus were investigated as a function of the deoxyglucose concentration. The apparent Km for deoxyglucose transport was about the same for all types of cells (1–2 mM). The Vmax of secondary cultures of mouse embryo cells decreased from 6 nmoles/106 cells/minute for sparse cultures to less than 1 nmole/106 cells/minute for density-inhibited cultures. The Vmax was about the same whether estimated in monolayer culture or in suspensions of cells dispersed by treatment with trypsin. The Vmax for deoxyglucose transport by the established cells, whether transformed spontaneously or by virus infection, was 4 to 25 times higher than that for density-inhibited mouse embryo cells and was independent of the cell density of the cultures. Deoxyglucose transport was competitively inhibited by Cytochalasin B, Persantin, glucose and 3-O-methyl-D-glucose and the apparent Ki values of inhibition were similar for the mouse embryo cells and the various cell lines. Similarly, the sensitivity of the glucose transport systems to inactivation by p-chloromercuribenzoate was about the same for all types of cells. The results suggest that the glucose transport system of the normal mouse embryo cells and the cells of the various established lines is qualitatively the same, but that the number of functional transport sites differs for the various cell lines and decreases markedly in mouse embryo cells with an increase in cell density of the cultures.  相似文献   

11.
The growth of transformed mouse fibroblasts (3T6 cells) in medium containing 5% fetal bovine serum was inhibited after treatment with concentrations greater than 50 microM ATP, ADP, or AMP. Adenosine, the common catabolite of the nucleotides, had no effect on cell growth at concentrations below 1 mM. However, the following results indicate that the toxicity of ATP, ADP, and AMP is mediated by serum- and cell-associated hydrolysis of the nucleotides to adenosine. 1) ADP and AMP, but not ATP, were toxic to 3T6 cells grown in serum-free medium or medium in which phosphohydrolase activity of serum was inactivated. Under these conditions, the cells exhibited cell-associated ADPase and 5'-nucleotidase activity, but little ecto-ATPase activity. 2) Inhibition of adenosine transport in 3T6 cells by dipyridamole or S-(p-nitrobenzyl)-6-thioinosine prevented the toxicity of ATP in serum-containing medium and of ADP and AMP in serum-free medium. 3) A 16-24-h exposure to 125 microM AMP or ATP was needed to inhibit cell growth under conditions where serum- and cell-associated hydrolysis of the nucleotides generated adenosine in the medium continuously over the same time period. In contrast, 125 microM adenosine was completely degraded to inosine and hypoxanthine within 8-10 h. Furthermore, multiple doses of adenosine added to the cells at regular intervals over a 16-h period were significantly more toxic than an equivalent amount of adenosine added in one dose. Treatment of 3T6 cells with AMP elevated intracellular ATP and ADP levels and reduced intracellular UTP levels, effects which were inhibited by extracellular uridine. Uridine also prevented growth inhibition by ATP, ADP, and AMP. These and other results indicate that serum- and cell-associated hydrolysis of adenine nucleotides to adenosine suppresses growth by adenosine-dependent pyrimidine starvation.  相似文献   

12.
From a mutagenized population of S49 murine T lymphoma cells, a mutant cell line, JPA4, was selected that expressed an altered nucleoside transport capability. JPA4 cells transported low concentrations of purine nucleosides and uridine more rapidly than the parental S49 cell line. The transport of these nucleosides by mutant cells was insensitive to inhibition by either dipyridamole (DPA) or 4-nitrobenzylthioinosine (NBMPR), two potent inhibitors of nucleoside transport in mammalian cells. Kinetic analyses revealed that the apparent Km values for the transport of uridine, adenosine, and inosine were 3-4-fold lower in JPA4 cells compared to wild type cells. In contrast, the transport of both thymidine and cytidine by JPA4 cells was similar to that of parental cells, and transport of these pyrimidine nucleosides remained sensitive to inhibition by both NBMPR and DPA. Furthermore, thymidine was a 10-12-fold weaker inhibitor of inosine transport in JPA4 cells than in wild type cells. Thus, JPA4 cells appeared to express two types of nucleoside transport activities; a novel (mutant) type that was insensitive to inhibition by DPA and NBMPR and transported purine nucleosides and uridine, and a parental type that retained sensitivity to inhibitors and transported cytidine and thymidine. The phenotype of the JPA4 cell line suggests that the sensitivity of mammalian nucleoside transporters to both NBMPR and DPA can be genetically uncoupled from its ability to transport certain nucleoside substrates and that the determinants on the nucleoside transporter that interact with each nucleoside are not necessarily identical.  相似文献   

13.
Time courses of [3H]uridine uptake as a function of uridine concentration were determined at 25° in untreated and ATP-depleted wild-type and uridine kinase-deficient Novikoff cells and in mouse L and P388 cells, Chinese hamster ovary cells and human HeLa cells. Short term uptake was measured by a rapid sampling technique which allows sampling of cell suspensions in intervals as short as one and one-half seconds. The initial segments of the time courses were the same in untreated, wild-type cells in which uridine is rapidly phosphorylated and in cells in which uridine phosphorylation was prevented due to lack of ATP or uridine kinase. The initial rates of uptake, therefore, reflected the rate of uridine transport. Uridine uptake, however, was approximately linear for only five to ten seconds at uridine concentrations from 20–160 μM and somewhat longer at higher concentrations. In phosphorylating cells the rate of uridine uptake (at 80 μM) then decreased to about 20–30% of the initial rate and this rate was largely determined by the rate of phosphorylation rather than transport. At uridine concentrations below 1 μM, however, the rate of intracellular phosphorylation in Novikoff cells approached the transport rate. The apparent substrate saturation of phosphorylation suggests the presence of a low Km uridine phosphorylation system in these cells. The “zero-trans” (zt) Km for the facilitated transport of uridine as estimated from initial uptake rates fell between 50 and 240 μM for all cell lines examined. The zero-trans Vmax values were also similar for all the lines (4–15 pmoles/μ1 cell H2O.sec). The time courses of uridine uptake by CHO cells and the kinetic constants for transport were about the same whether the cells were propagated (and analyzed for uridine uptake) in suspension or monolayer culture. When Novikoff cells were preloaded with 10 μM uridine the apparent Km and Vmax values (infinite-trans) were two to three times higher than the corresponding zero-trans values. Uridine transport was inhibited in a simple competitive manner by several other ribo- and deoxyribonucleosides. All nucleosides seem to be transported by the same system, but with different efficiencies. Uridine transport was also inhibited by hypoxanthine, adenine, thymine, Persantin, papaverin, and o-nitrobenzylthioinosine, and by pretreatment of the cells with p-chloromercuri-benzoate, but not by high concentrations of cytosine, D-ribose or acronycin. The inhibition of uridine transport by Persantin involved changes in both V and K. Because of the rapidity of transport, some loss of intracellular uridine occurred when cells were rinsed in buffer solution to remove extracellular substrate, even at 0°. This loss was prevented by the presence of a transport inhibitor, Persantin, in the rinse fluid or by separating suspended cells from the medium by centrifugation through oil. Metabolic conversion of intracellular uridine were also found to continue during the rinse period. The extent of artifacts due to efflux and metabolism during rinsing increased with duration of the rinse.  相似文献   

14.
The zero-trans influx of 500 microM uridine by CHO, P388, L1210 and L929 cells was inhibited by nitrobenzylthioinosine ( NBTI ) in a biphasic manner; 60-70% of total uridine influx by CHO cells and about 90% of that in P388, L1210 and L929 cells was inhibited by nmolar concentrations of NBTI (ID50 = 3-10 nM) and is designated NBTI -sensitive transport. The residual transport activity, designated NBTI -resistant transport, was inhibited by NBTI only at concentrations above 1 microM (ID50 = 10-50 microM). S49 cells exhibited only NBTI -sensitive uridine transport, whereas Novikoff cells exhibited only NBTI -resistant uridine transport. In all instances NBTI -sensitive transport correlated with the presence of between 7 7 X 10(4) and 7 X 10(5) high-affinity NBTI binding sites/cell (Kd = 0.3-1 nM). Novikoff cells lacked such sites. The two types of nucleoside transport, NBTI -resistant and NBTI -sensitive, were indistinguishable in substrate affinity, temperature dependence, substrate specificity, inhibition by structurally unrelated substances, such as dipyridamole or papaverine, and inhibition by sulfhydryl reagents or hypoxanthine. We suggest, therefore, that a single nucleoside transporter can exist in an NBTI -sensitive and an NBTI -resistant form depending on its disposition in the plasma membrane. The sensitive form expresses a high-affinity NBTI binding site(s) which is probably made up of the substrate binding site plus a hydrophobic region which interacts with the lipophilic nitrobenzyl group of NBTI . The latter site seems to be unavailable in NBTI -resistant transporters. The proportion of NBTI -resistant and sensitive uridine transport was constant during proportion of NBTI -resistant and sensitive uridine transport was constant during progression of P388 cells through the cell cycle and independent of the growth stage of the cells in culture. There were additional differences in uridine transport between cell lines which, however, did not correlate with NBTI sensitivity and might be related to the species origin of the cells. Uridine transport in Novikoff cells was more sensitive to inhibition by dipyridamole and papaverine than that in all other cell lines tested, whereas uridine transport in CHO cells was the most sensitive to inactivation by sulfhydryl reagents.  相似文献   

15.
Activated hepatic stellate cells (HSC; lipocytes; Ito cells) proliferate and are responsible for extracellular matrix synthesis during hepatic fibrogenesis. During activation, HSC undergo transdifferentiation into myofibroblasts expressing alpha-smooth muscle actin (alpha-SMA). Adenosine 3', 5'-cyclic monophosphate (cyclic AMP) is an ubiquitous intracellular signaling molecule, and is upregulated by the activation of adenylate cyclase and downregulated via hydrolysis by cyclic nucleotide phosphodiesterases (PDEs). Recently, increased intracellular cyclic AMP has been shown to inhibit HSC activation. The aim of the current study was to determine the effects of inhibition of PDEs on cell proliferation and transdifferentiation in cultured rat HSC. Cell proliferation was determined by [3H]thymidine incorporation, and Western blot analysis was performed for detection of alpha-SMA, a phenotypic marker of transdifferentiation into myofibroblast. When the cells were exposed to 3-isobutyl-1-methylxanthine (IBMX; 50-1000 microM), a nonselective PDE inhibitor, serum-stimulated [3H]thymidine incorporation was suppressed in a dose-dependent manner with a maximum inhibition of 66% at a concentration of 500 microM OPC-13013 (1-60 microM), a selective PDE III isoenzyme inhibitor, induced a dose-dependent inhibitory effect on serum-stimulated DNA synthesis that reached a maximum inhibition of 95% at a concentration of 60 microM, while neither 8-methoxymethyl-3-isobutyl-1-methylxanthine (8-MMX), a PDE I isoenzyme inhibitor, nor Ro-20-1724, a PDE IV isoenzyme inhibitor, had an inhibitory effect. Western blot analysis revealed that IBMX or OPC-13013 decreased alpha-SMA expression, while other selective PDE isoenzyme inhibitors did not have a suppressive effect. IBMX, OPC-13013 or Ro-20-1724, but not 8-MMX augmented forskolin-induced increase in intracellular cyclic AMP levels although cyclic AMP levels were not affected by treatment with any of these PDE inhibitors alone. These data indicate that inhibition of PDEs, especially PDE III isoenzyme, can produce an inhibitory effect on HSC activation. The PDE III isoenzyme may contribute to the regulation of HSC activation during fibrogenesis. In addition, OPC-13013 may have the potential to inhibit initiation and progression of hepatic fibrosis by interfering with HSC activation.  相似文献   

16.
N62-Isopentenyl)adenosine (IPAR) inhibited severely the incorporation of uridine and cytidine into S-180 cells in culture. When IPAR and the nucleosides were simultaneously present in the medium the inhibition was competitive (Ki 3.4 m̈M) and indicated inhibition of transport. However, the inhibition occurred even in the absence of extracellular IPAR if the cells had been preincubated with IPAR. Since 5′-IPAMP was the product which accumulated in large quantities in S-180 cells when incubated with IPAR, the effects of this AMP analog on the intracellular metabolism of uridine had to be considered. No direct correlation between the amount of intracellular IPAMP and the degree of inhibition of uridine utilization was observed and the relative distribution of uridine nucleotides in the acid soluble pool of the cells was unaltered in cells treated with IPAR. Also, IPAMP was not an inhibitor of uridine kinase in a cell free system nor was the activity of this enzyme affected by treatment of cells with IPAR. In addition, a profound inhibition of uridine utilization was also observed in a resistant subline of S-180 cells, which is unable to form IPAMP. These data suggest that IPAMP was not the inhibitory agent. Furthermore, the observation that the inhibition in both sensitive and resistant cells was caused even by a 15-second exposure to 100 m̈M IPAR, followed by rinsing, suggests that IPAR itself is the effective agent. It is concluded that IPAR exerts its inhibitory effect on uridine and cytidine utilization by becoming lodged in the cell membrane and thereby preventing the passage of these nucleosides into the cells. It is also shown that the inhibition of uridine and cytidine utilization by IPAR and by other potent nucleoside uptake inhibitors is unrelated to inhibition of growth or of RNA-synthesis when the cells do not depend on an extracellular source of a nucleoside for growth.  相似文献   

17.
S49 mouse lymphoma cells are deficient in hypoxanthine transport   总被引:1,自引:0,他引:1  
The rate of uptake of hypoxanthine in S49 cells was only about 2-5% of the rate of hypoxanthine transport observed in many other types of mammalian cells, and of the rate of uridine transport in this and other cell types. Part of the slow entry of hypoxanthine seems to be due to non-mediated permeation, but the remainder is saturable, strongly inhibited by uridine, nitrobenzylthioinosine and dipyridamole and not detectable in a nucleoside-transport-deficient mutant of S49 cells (AE1). The inhibition of hypoxanthine transport in S49 cells by nitrobenzylthioinosine resembles the inhibition of nucleoside transport in these and other mammalian cells, whereas it contrasts with the resistance of hypoxanthine transport to nitrobenzylthioinosine in all types of mammalian cells that have been investigated. We conclude that S49 cells lack the hypoxanthine transport system common to other types of cells and that hypoxanthine entry into these cells is mediated, although very inefficiently, by the nucleoside transporter. In contrast, adenine transport in S49 and AE1 cells was comparable to that in other types of cells.  相似文献   

18.
Rapid kinetic techniques were used to study the transport and salvage of uridine and other nucleosides in mouse spleen cells. Spleen cells express two nucleoside transport systems: (1) the non-concentrative, symmetrical, Na+-independent transporter with broad substrate specificity, which has been found in all mammalian cells and is sensitive to inhibition by dipyridamole and nitrobenzylthioinosine; and (2) a Na+-dependent nucleoside transport, which is specific for uridine and purine nucleosides and resistant to inhibition by dipyridamole and nitrobenzylthioinosine. The kinetic properties of the two transporters were determined by measuring uridine influx in ATP-depleted cells and dipyridamole-treated cells, respectively. The Michaelis-Menten constants for Na+-independent and -dependent transport were about 40 and 200 microM, respectively, but the first-order rate constants were about the same for both transport systems. Nitrobenzylthioinosine-sensitivity of the facilitated nucleoside transporter correlated with the presence of about 10,000 high-affinity (Kd = 0.6 nM) nitrobenzylthioinosine-binding sites per cell. The turnover number of the nitrobenzylthioinosine-sensitive nucleoside transporter was comparable to that of mouse P388 leukemia cells. The activation energy of this transporter was 20 kcal/mol. Entry of uridine via either of the transport routes was rapidly followed by its phosphorylation and conversion to UTP. The Michaelis-Menten constant for the in situ phosphorylation of uridine was about 50 microM and the first-order rate constants for phosphorylation and transport were about the same. The spleen cells also efficiently salvaged adenosine, adenine, and hypoxanthine, but not thymidine.  相似文献   

19.
The Daudi line of human lymphoblastoid cells shows a high sensitivity towards growth inhibition by human interferons. In cells pretreated with 70 reference units/ml of an interferon preparation for 48 h, the incorporation of exogenous [3H]thymidine into DNA is inhibited by as much as 85%. We are investigating the extent to which this effect reflects a true inhibition of the rate of DNA synthesis or whether it may be caused by changes in the metabolic utilization of exogenous thymidine by the cells. Interferon treatment results in a 30% inhibition of the rate of membrane transport and a 60% decrease in the rate of phosphorylation of [3H]thymidine in vivo. The latter effect is due to a decrease in V of thymidine kinase without any change in the value of Km for this enzyme. In addition to these changes, incorporation of [3H]uridine into DNA, which occurs as a result of the intracellular conversion of this precursor into thymidine nucleotides, is also inhibited by 75%, whereas RNA labelling by [3H]uridine is decreased by only 15% in interferon-treated cells. Thus several different metabolic events associated with thymidine nucleotide metabolism and DNA synthesis in Daudi cells are disrupted by interferon treatment.  相似文献   

20.
The zero-trans influx of 500 μM uridine by CHO, P388, L1210 and L929 cells was inhibited by nitrobenzylthioinosine (NBTI) in a biphasic manner; 60–70% of total uridine influx by CHO cells and about 90% of that in P388, L1210 and L929 cells was inhibited by nmolar concentrations of NBTI (ID50 = 3?10 nM) and is designated NBTI-sensitive transport. The residual transport activity, designated NBTI-resistant transport, was inhibited by NBTI only at concentrations above 1 μM (ID50 = 10?50 μM). S49 cells exhibited only NBTI-sensitive uridine transport, whereas Novikoff cells exhibited only NBTI-resistant uridine transport. In all instances NBTI-sensitive transport correlated with the presence of between 7·104 and 7·105 high-affinity NBTI binding sites/cell (Kd = 0.3?1 nM). Novikoff cells lacked such sites. The two types of nucleoside transport, NBTI-resistant and NBTI-sensitive, were indistinguishable in substrate affinity, temperature dependence, substrate specificity, inhibition by structurally unrelated substances, such as dipyridamole or papaverine, and inhibition by sulfhydryl reagents or hypoxanthine. We suggest, therefore, that a single nucleoside transporter can exist in an NBTI-sensitive and an NBTI-resistant form depending on its disposition in the plasma membrane. The sensitive form expresses a high-affinity NBTI binding site(s) which is probably made up of the substrate binding site plus a hydrophobic region which interacts with the lipophilic nitrobenzyl group of NBTI. The latter site seems to be unavailable in NBTI-resistant transporters. The proportion of NBTI-resistant and sensitive uridine transport was constant during proportion of NBTI-resistant and sensitive uridine transport was constant during progression of P388 cells through the cell cycle and independent of the growth stage of the cells in culture. There were additional differences in uridine transport between cell lines which, however, did not correlate with NBTI sensitivity and might be related to the species origin of the cells. Uridine transport in Novikoff cells was more sensitive to inhibition by dipyridamole and papaverine than that in all other cell lines tested, whereas uridine transport in CHO cells was the most sensitive to inactivation by sulfhydryl reagents.  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