首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
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   

2.
Summary The sensitivity of nucleoside transport by rat erythrocytes to inhibition by nitrobenzylthioinosine (NBMPR) and the slowly permeating organomercurial,p-chloromercuriphenyl sulfonate (pCMBS), was investigated. The dose response curve for the inhibition of uridine transport (100 M) by NBMPR was biphasic –35% of the transport activity was inhibited with an IC50 value of 0.25 nM, but 65% of the activity remained insensitive to concentrations as high as 1 M. These two components of uridine transport are defined as NBMPR-sensitive and NBMPR-insensitive, respectively. Uridine influx by both components was saturable and conformed to simple Michaelis-Menten kinetics, and was inhibited by other nucleosides. The uridine affinity of the NBMPR-sensitive transport component was threefold higher than for the NBMPR-insensitive transport mechanism (apparentK m for uridine 50±18 and 163±28 M, respectively). The two transport systems also differed in their sensitivity topCMBS. NBMPR-insensitive uridine transport was inhibited bypCMBS with an IC50 of 25M, while 1 mMpCMBS had little effect on NBMPR-sensitive transport by intact cells.pCMBS inhibition was reduced in the presence of uridine and adenosine and reversed by the addition by -mercaptoethanol, suggesting that thepCMBS-sensitive thiol group is located on the exterior surface of the erythrocyte membrane within the nucleoside binding site of the transport system. Inhibition of uridine transport by NBMPR was associated with high-affinity [3H]NBMPR binding to the cell membrane (apparentK d46±25 pM). Binding of inhibitor to these sites was competitively blocked by uridine and inhibited by adenosine, thymidine, dipyridamole, dilazep and nitrobenzylthioguanosine. Assuming that each NBMPR-sensitive transport site binds a single molecule of NBMPR, the calculated translocation capacity of each site is 25±6 molecules/site per sec at 22°C.pCMBS had no effect on [3H]NBMPR binding to intact cells but markedly inhibited binding to disrupted membranes indicating that the NBMPR-sensitive nucleoside transporter probably has a thiol group located on the inner surface of the membrane. Exposure of rat erythrocyte membranes to UV light in the presence of [3H]NBMPR resulted in covalent radiolabeling of a membrane protein(s) (apparent Mr on SDS gel electropherograms of 62,000). Labeling of this protein was abolished in the presence of nitrobenzylthioguanosine. We conclude that nucleoside transport by rat erythrocytes occurs by two facilitated-diffusion systems which differ in their sensitivity to inhibition by both NBMPR andpCMBS.  相似文献   

3.
The kinetic mechanism of chlorpromazine inhibition of erythrocyte hexose transport was investigated using the non-metabolizable glucose analog 3-O-methylglucose. It was found that chlorpromazine added to the external medium is a non-competitive inhibitor of both equilibrium exchange and net 3-O-methylglucose transport at pH 7.8, 15°C. The Ki for equilibrium exchange is 76 ± 21 μM. When net efflux and equilibrium exchange were measured on the same population of cells the equilibrium exchange was 2.5-times the maximum net efflux. The percent reduction of 3-O-methylglucose flux by chlorpromazine is dependent upon chlorpromazine concentration and not 3-O-methylglucose concentration as expected for a non-competitive inhibitor. Equilibrium exchange and net efflux show the same extent of inhibition at each concentration of chlorpromazine evaluated. These results suggest that exchange and net efflux of 3-O-methylglucose in the human erythrocyte may share a common transport system.  相似文献   

4.
The rate of exchange of HCO3? with Cl? at 37°C in erythrocytes of ten mammalian species was studied. The rate constant increases from 7s?1 (ox) to 16s?1 (rat), and is inversely proportional to the body size (log10) of the species. It is found that the membrane permeability in different species is positively correlated to the red cell membrane phosphatidylcholine or arachidonate content, and is negatively correlated to the sphingomyelin or linoleate content.  相似文献   

5.
The transport of thymidine has been characterized kinetically and thermodynamically in Novikoff rat hepatoma cells grown in culture and, less extensively, in mouse L cells, Chinese hamster ovary cells, P388 murine leukemia cells and HeLa cells. That the characterizations pertained to the transport system per se was ensured, (i) by employing recently developed methods for rapid sampling of cell/substrate mixtures in order to follow isotope movements within a few seconds after initial exposure of cells to substrate; (ii) by utilizing cells rendered, by genetic or chemical means, incapable of metabolizing thymidine; and, (iii) by demonstrating conformity of the transport data to an integrated rate equation derived for a simple, carrier-mediated system. The results indicate that thymidine is transported into mammalian cells by a functionally symmetrical, non-concentrative system for which the carrier : substrate dissociation constant ranges from about 100 μM in Chinese hamster ovary cells, to 230 μM in Novikoff hepatoma cells. In all cell lines investigated, the velocity of transport was sufficient to nearly completely equilibrate low concentrations of thymidine across the membrane within 15 s. Temperature dependence of transport velocity and substrate : carrier dissociation were continuous (EA = 18.3 kcal/mol, ΔH0′ = 9.3 kcal/mol, respectively), and showed no evidence of abrupt transitions. Several natural and artificial nucleosides and nucleic acid based inhibited influx of radiolabeled thymidine, apparently by competing with thymidine for the transport carrier.  相似文献   

6.
Arginine transport in suspension-cultured cells of Nicotiana tabacum L. cv. Wisconsin-38 was investigated. Cells that were preincubated in the presence of Ca2+ for 6 h prior to transport exhibited stimulated transport rates. After the preincubation treatment, initial rates of uptake were constant for at least 45 min. Arginine accumulated in the cells against a concentration gradient; this accumulation was not the result of exchange diffusion. Arginine uptake over a concentration range of 2.5 M to 1 mM was characterized by simple Michaelis-Menten kinetics with a Km of 0.1 mM and a Vmax of 9,000 nmol g-1 fresh weight h-1. Transport was inhibited by several compounds including carbonylcyanide-m-chlorophenylhydrazone, 2,4-dinitrophenol, N,N-dicyclohexylcarbodiimide, and N-ethylmaleimide. Inhibition by these compounds was not the result of increased efflux resulting from membrane damage. A variety of amino acids and analogs, with the exception of D-arginine, inhibited transport, indicating that arginine transport was mediated by a general L-aminoacid permease. Competition experiments indicated that arginine and lysine exhibited cross-competition for transport, with Ki values similar to respective Km values. Arginine transport and low-affinity lysine transport are probably mediated by the same system in these cells.Abbreviations BTP Bis Tris Propane - CCCP Carbonylcyanide-m-chlorophenylhydrazone - DCCD N,N-dicyclohexylcarbodiimide - DNP 2,4-dinitrophenol - DTT Dithiothreitol - NEM N-ethylmaleimide - MES 2(N-morpholino)ethanesulfonic acid - TCA trichloroacetic acid This paper is the third in a series on amino-acid transport into cultured tobacco cells. For parts I and II, see Harrington and Henke (1981) and Harrington et al. (1981)  相似文献   

7.
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   

8.
(1) Unidirectional K+ (86Rb) influx and efflux were measured in subconfluent layers of MDCK renal epithelial cells and HeLa carcinoma cells. (2) In both MDCK and HeLa cells, the furosemide-inhibitable and chloride-dependent component of K+ influx/efflux was stimulated 2-fold by a 30 min incubation in 1 · 10?3 M ouabain. (3) Measurements of net K+ loss and Na+ gain in ouabain-treated cells at 1 h failed to show any diuretic sensitive component, confirming the exchange character of the diuretic-sensitive fluxes. (4) Prolonged incubations for 2.5 h in ouabain revealed a furosemide- and anion-dependent K+ (Cl?) outward net flux uncoupled from net Na+ movement. Net K+ (Cl?) outward flux was half-maximally inhibited by 2 μM furosemide. (5) After 2.5 h ouabain treatment, the anion and cation dependence of the diuretic-sensitive K+ influx/efflux were essentially unchanged when compared to untreated controls.  相似文献   

9.
Methotrexate, 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide and N-hydroxysuccinimide react to form an activated ester of methotrexate which is a potent irreversible inhibitor of methotrexate transport in L1210 cells. In cells treated with the reagent at 37°C, inhibition was rapid (t12 < 1 min), optimal at pH 6.8, half-maximal at an inhibitor concentration of 20 nM, and complete at high levels of the reagent. Specificity was indicated by the fact that excess methotrexate added during the pretreatment step protected the transport system against inactivation. Irreversible inhibition was also observed in cells exposed to the reagent at 4°C. Inactivation in this case was qualitatively similar to the corresponding process at 37°C; it appeared rapidly, was half-maximal at 20 nM, and could be prevented by the addition of high concentrations of the substrate. The extent of the inhibition, however, reached a maximum of only 75%, even in samples containing excess or multiple additions of reagent. The latter findings suggest that at 4°C the transport protein exists in two forms, one (75% of the total) containing binding sites which are accessible to the active ester, and the other (25% of the total) with inaccessible sites. The identity of these sites is suggested to be transport proteins which have outward and inward orientations, respectively.  相似文献   

10.
11.
Cells of an adenosine-resistant clone (AE1) of S49 mouse lymphoma cells were compared with cells of the parental line with respect to (a) characteristics of nucleoside transport, (b) high affinity binding of the inhibitor of nucleoside transport, nitrobenzylthionisine (NBMPR), and (c) the antiproliferative effects of the nucleoside antibiotics, tubercidin, arabinosyladenine and showdomycin. Rates of inward transport of uridine, thymidine, adenosine, 2′-deoxyadenosine, tubercidin, showdomycin, and arabinosyladenine in AE1 cells were less than 1% of those in cells of the parental S49 line. The inhibitor of nucleoside transport, NBMPR, reduced rates of inward nucleoside transport in S49 cells to levels comparable to those seen in the transport-defective mutant. S49 cells possessed high affinity sites that bound NBMPR (6.6 · 104 sites/cell, Kd  0.2 nM), whereas site-specific binding of NBMPR to AE1 cells was not demonstrable, indicating that loss of nucleoside transport activity in AE1 cells was accompanied by loss of the high affinity NBMPR binding sites. Relative to S49 cells, AE1 cells were resistant to the antiproliferative effects of tubercidin and showdomycin, but differences between the two cell lines in sensitivity toward arabinosyladenine were minor, suggesting that nucleoside transport activity was required for cytotoxicity of tubercidin and showdomycin, but not for that of arabinosyladenine.  相似文献   

12.
Furosemide (1 · 10?4M) inhibits a proportion of the total passive (ouabain-insensitive) K+ influx into primary chick heart cell cultures (85%), BC3H1 cells (75%), MDCK cells (40%) and HeLa cells (57%). This action of furosemide upon K+ influx is independent of (Na+ + K+)-pump inhibition since the furosemide-sensitive component of the K+ influx is identical in the presence and absence of ouabain (1 · 10?3M). For HeLa cells the passive, furosemide-sensitive component of K+ influx is markedly dependent upon the external K+, Na+ and Cl? content. Acetate, iodide and nitrate are ineffective as substitutes for Cl?, whereas Br? is partially effective. Partial Cl? replacement by NO3? gave an apparent affinity of 100 mM [Cl]. Na+ replacement by choline+ abolishes the furosemide-sensitive component, whereas Li+ replacement reduces this component by 48%. Partial Na+ replacement by choline+ gives an apparent affinity of 25 mM [Na+]. Variation in the external K+ content gives an affinity for the furosemide-sensitive component of approx. 1.0 mM. Furosemide inhibition of the passive K+ inflúx is of high affinity, half-maximal inhibition being observed at 5 · 10?6M furosemide. Piretanide (1 · 10?4M) and phloretin (1 · 10?4M) inhibit the same component of passive K+ influx as furosemide; ethacrynic acid and amiloride (both 1 · 10?4M) partially so. The stilbene, SITS (1 · 10?6M), was ineffective as an inhibitor of the furosemide-sensitive component.  相似文献   

13.
The transport of [14C]phenylacetic acid (PAA) in intact plants and stem segments of light-grown pea (Pisum sativum L. cv. Alderman) plants was investigated and compared with the transport of [14C]indiol-3yl-acetic acid (IAA). Although PAA was readily taken up by apical tissues, unlike IAA it did not undergo long-distance transport in the stem. The absence of PAA export from the apex was shown not to be the consequence of its failure to be taken up or of its metabolism. Only a weak diffusive movement of PAA was observed in isolated stem segments which readily transported IAA. When [1-14C]PAA was applied to a mature foliage leaf in light, only 5.4% of the 14C recovered in ethanol extracts (89.6% of applied 14C) had been exported from the leaf after 6.0 h. When applied to the corresponding leaf, [14C]sucrose was readily exported (46.4% of the total recovered ethanol-soluble 14C after 6.0 h). [1-14C]phenylacetic acid applied to the root system was readily taken up but, after 5.0 h, 99.3% of the recovered 14C was still in the root system.When applied to the stem of intact plants (either in lanolin at 10 mg·g-1, or as a 10-4 M solution), unlabelled PAA blocked the transport through the stem of [1-14C]IAA applied to the apical bud, and caused IAA to accumulate in the PAA-treated region of the stem. Applications of PAA to the stem also inhibited the basipetal polar transport of [1-14C]IAA in isolated stem segments. These results are consistent with recent observations (C.F. Johnson and D.A. Morris, 1987, Planta 172, 400–407) that no carriers for PAA occur in the plasma membrane of the light-grown pea stem, but that PAA can inhibit the carrier-mediated efflux of IAA from cells. The possible functions of endogenous PAA are discussed and its is suggested that an important role of the compound may be to modulate the polar transport and-or accumulation by cells of IAA.Abbreviations IAA indol-3yl-acetic acid - NPA N-1-naphthylphthalamic acid - PAA phenylacetic acid - IIBA 2,3,5-triiodobenzoic acid  相似文献   

14.
Mitochondrial DNA depletion caused by thymidine kinase 2 (TK2) deficiency can be compensated by a nucleoside kinase from Drosophila melanogaster (Dm-dNK) in mice. We show that transgene expression of Dm-dNK in Tk2 knock-out (Tk2−/−) mice extended the life span of Tk2−/− mice from 3 weeks to at least 20 months. The Dm-dNK+/−Tk2−/− mice maintained normal mitochondrial DNA levels throughout the observation time. A significant difference in total body weight due to the reduction of subcutaneous and visceral fat in the Dm-dNK+/−Tk2−/− mice was the only visible difference compared with control mice. This indicates an effect on fat metabolism mediated through residual Tk2 deficiency because Dm-dNK expression was low in both liver and fat tissues. Dm-dNK expression led to increased dNTP pools and an increase in the catabolism of purine and pyrimidine nucleotides but these alterations did not apparently affect the mice during the 20 months of observation. In conclusion, Dm-dNK expression in the cell nucleus expanded the total dNTP pools to levels required for efficient mitochondrial DNA synthesis, thereby compensated the Tk2 deficiency, during a normal life span of the mice. The Dm-dNK+/− mouse serves as a model for nucleoside gene or enzyme substitutions, nucleotide imbalances, and dNTP alterations in different tissues.  相似文献   

15.
Methotrexate transport in L1210 cells is highly sensitive to inhibition by p-chloromercuriphenylsulfonate (CMPS) and, to a lesser extent, by N-ethylmaleimide. A 50% reduction in the methotrexate influx rate occurred upon exposure of cells to 3 μM CMPS or 175 μM N-ethylmaleimide, while complete inhibition was achieved at higher levels of these agents. Dithiothreitol reversed the inhibition by CMPS, suggesting that a sulfhydryl residue is involved. This residue is apparently not located at the substrate binding site of the transport protein, since methotrexate failed to protect the system from inactivation by either CMPS or N-ethylmaleimide, and the transport protein retained the ability to bind substrate (at 4°C) after exposure to these inhibitors (at 37°C). Methotrexate efflux was also inhibited by CMPS (50% at 4 μM), indicating that both the uptake and efflux of methotrexate in L1210 cells occur via the same transport system. High concentrations of CMPS (greater than 20 μM) increased the efflux rate, apparently by damaging the cell membrane and allowing the passive diffusion of methotrexate out of the cell.  相似文献   

16.
External ATP causes a passive permeability change in several transformed cells, but not in untransformed cells. We previously demonstrated that in CHO-K1 cells, a transformed clone of Chinese hamster ovary cells, the external ATP-dependent permeability change was induced when the intracellular ATP concentration was reduced by a mitochondrial inhibitor (Kitagawa, T. and Akamatsu, Y. (1981) Biochim. Biophys. Acta 649, 76–82). A permeability change with similar characteristics was also observed when the CHO cells were treated with external ATP and a cytoskeleton-attacking agent such as vinblastine or cytochalasin B. Just like mitochondrial inhibitors, vinblastine could increase the sensitivity of transformed 3T3 cells to external ATP but showed no effect on passive permeability of normal 3T3 cells. However, in contrast with the effect of the mitochondrial inhibitors, the cytoskeleton drugs caused the permeability change with little reduction of intracellular ATP concentration, suggesting different actions of these two kinds of drug on the permeability change. The present results suggest an important role of cytoskeletal structures in controlling the external ATP-dependent permeability change in transformed cells. Possible effects of intracellular ATP on cytoskeletal structures are also discussed.  相似文献   

17.
Embryonic chick heart cells in culture transport 2-deoxy-D-glucose and 3-O-methyl-D-glucose very rapidly. By direct measurements of uptake, it was not possible to estimate accurately transport rates, nor, with 2-deoxyglucose, to discriminate clearly between its transport and phosphorylation. In contrast, the technique of countertransport made it possible to determine precisely initial transport velocity and to make the following observations: (1) phosphorylation, and not transport, is rate-limiting in 2-deoxyglucose uptake; (2) hexose transport is stimulated 5-fold by removal of glucose from culture medium; and (3) this stimulation is followed by an increase in phosphorylation, but the effect is much less pronounced (2-fold stimulation only). In conclusion, the adaptative regulation of glucose transport described in many fibroblast cell lines exists also in cardiac cells.  相似文献   

18.
Both enzyme-mediated group translocation and facilitated diffusion have been proposed as mechanisms by which mammalian cells take up purine bases and nucleosides. We have investigated the mechanisms for hypoxanthine and inosine transport by using membrane vesicles from Chinese hamster ovary cells (CHO), Balb/c 3T3 and SV3T3 cells prepared by identical procedures. Uptake mechanisms were characterized by analyzing intravesicular contents, determining which substrates could exchange with the transport products, assaying for hypoxanthine phosphoribosyltransferase activity, and measuring the stimulation of uptake of hypoxanthine by phosphoribosyl pyrophosphate (PRib-PP).We found that the uptake of hypoxanthine in Balb 3T3 vesicles was stimulated 3–4-fold by PRib-PP. The intravesicular product was predominantly IMP. The hypoxanthine phosphoribosyltransferase activity copurified with the vesicle preparation. These results suggest the possible involvement of this enzyme in hypoxanthine uptake in 3T3 vesicles. In contrast to the 3T3 vesicles, CHO vesicles prepared under identical procedures did not retain hypoxanthine phosphoribosyltransferase activity and did not demonstrate PRib-PP-stimulated hypoxanthine uptake. The intravesicular product of hypoxanthine uptake in CHO vesicles was hypoxanthine. These results and data from our kinetic and exchange studies indicated that CHO vesicles transport hypoxanthine via facilitated diffusion. An analogous situation was observed for inosine uptake; CHO vesicles accumulated inosine via a facilitated diffusion mechanism, while in the same experiments SV3T3 vesicles exhibited a purine nucleoside phosphorylase-dependent translocation of the ribose moiety of inosine.  相似文献   

19.
Depletion of energy stores of human red cells decreases the maximum transport capacity, Jm, for glucose transport to a value one-third or less of that found in red cells from freshly drawn blood. There is no change in Km. Hemolysis and resealing of red cells with ATP or ADP reverses the decrease in Jm. The maximum effect occurs at concentrations of ATP in the normal range for red cells, however, there is little effect from ADP concentrations in its normal range in freshly drawn red cells. Hemolysis and resealing with ATP gives an increase in Jm and an increase in differential labeling by photolytic labeling with tritiated cytochalasin B. Most of the activation is lost after a second hemolysis-reseal without ATP but about 25% of the activation remains.  相似文献   

20.
F Pelaez  C de Haro 《FEBS letters》1989,250(2):523-528
The polypeptide chain initiation factor 2 (eIF-2) binds phospholipid (PL) and becomes a potent inhibitor of translation in hemin-supplemented reticulocyte lysates [De Haro et al. (1986) Proc. Natl. Acad. Sci. USA 83, 6711–6715]. This binding is independent of calcium ions and seems to be specific for phosphatidylinositol or phosphatidylserine; phosphatidic and arachidonic acids are inactive. Like -subunit-phosphorylated eIF-2, eIF-2·PL traps GEF in a non-dissociable eIF-2·PL·GEF complex whereby GEF is no longer able to recycle. Initiation is inhibited when no free GEF is available. Translational inhibition by eIF-2·PL is rescued by equimolar amounts of eIF-2·GEF. On the basis of this stoichiometry, we have estimated that reticulocyte lysates contain about 60 pmol of GEF/ml (60 nM) eIF-2·PL also inhibits translation in cell-free mouse liver extracts and this inhibition is prevented by reticulocyte eIF-2·GEF suggesting that GEF also functions in liver. However, the eIF-2·PL complex does not affect translation in such non-mammalian eukaryotic systems as wheat germ and Drosophila embryos.  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