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.
A problem typically encountered in the analysis of amino acids in chemical evolution experiments and in extracts of meteorites is the large number present. For example, α-, β-, and γ-amino acids, N-mono substituted α-amino acids, and dicarboxylic α-amino acids have been found in extracts of the Murchison meteorite, and many more amino acids are present than have been positively identified by computerized gas chromatographic mass spectrometry. This paper reports an analytical method to selectively destroy the α-amino acids, with only the β- and γ-amino acids remaining in the solution. It is based on the ability of Cu2+ to complex with amino acids, the order of stability of these complexes being α > β > γ, = δ, = ε = 0. Aqueous solutions of α-amino acid-Cu2+ chelates are known to be decomposed by 254 nm light as well as by nonmonochromatic uv light, yielding a precipitate of Cu2O. This paper shows that at 254 nm (ligand-metal charge transfer band) the rate of destruction of amino acids in Cu2+ aqueous solutions is in the following order, dicarboxylic α-amino acids > α-amino acids > N-monosubstituted α-amino acids β-amino acids ≈ γ-amino acids. Thus by irradiation with 254 nm light in the presence of Cu2+ all the amino acids can be destroyed except the β- and γ-amino acids. When almost 100% of the α-amino acids are destroyed, 80% of the β- and γ-amino acids still exist in solution. With this procedure, complex mixtures of amino acids can be simplified to make identification by gas chromatographic mass spectrometry casier.  相似文献   

2.
The serine proteinase α-thrombin potently stimulates reinitiation of DNA synthesis in quiescent Chinese hamster fibroblasts (CCL39 line). 125I-labeled α-thrombin binds rapidly and specifically to CCL39 cells with high affinity (Kd ≈ 4 nM). Binding at 37°C was found to remain stable for 6 h or more during which time no receptor down-regulation, ligand internalization and/or degradation could be detected. The structure of α-thrombin receptors on CCL39 cells was identified by covalently coupling 125I-α-thrombin to intact cells using a homobifunctional cross-linking agent, ethylene glycol bis(succinimidyl succinate). By resolution in sodium dodecyl sulfate polyacrylamide gel electrophoresis we observed the specific labeling of a major α-thrombin-binding site of Mr ≈ 150 000 revealed as a 125I-α-thrombin cross-linked complex of Mr ≈ 180 000. Independent of chemical cross-linking, 125I-α-thrombin also formed a covalent complex with a minor, 35 000 Mr, membrane component identified as protease nexin. Two derivatives of α-thrombin modified at the active site are 1000-fold less than α-thrombin for mitogenicity. These two non-mitogenic derivatives bound to cells with similar affinity and maximal binding capacity as native α-thrombin, and affinity-labeled the receptor subunit of Mr 150 000. When present in large excess, during incubation of cells with α-thrombin, these binding antagonists were ineffective in blocking α-thrombin-induced DNA synthesis. These data suggest that the specific 150 000 Mr binding sites that display high affinity for α-thrombin do not mediate induction of the cellular mitogenic response.  相似文献   

3.
A secretory granule-associated enzymatic activity that converts mono-[125I]-D-Tyr-Val-Gly into mono-[125I]-D-Tyr-Val-NH2 has been studied. The activity is primarily soluble and shows optimal activity at pH 7 to pH 8. Amidation activity was stimulated 9-fold by addition of optimal amounts of copper (3 μM). In the presence of optimal copper, ascorbate stimulated the reaction 7-fold; none of the other reduced or oxidized cofactors tested was as effective. Taking into account the dependence of the reaction on ascorbate and molecular oxygen and the production of glyoxylate [2], it is suggested that the α-amidation enzyme is a monooxygenase. Lineweaver Burk plots with D-Tyr-Val-Gly as the varied substrate demonstrated Michelis-Menten type kinetics with the values of Km and Vmax increasing with the addition of ascorbate to the assay. A variety of peptides ending with a COOH-terminal Gly residue act as inhibitors of the reaction. Two synthetic peptides, γ2MSH and ACTH(1–14), with carboxyl termini similar to the presumed physiological substrates for the enzyme, act as competitive inhibitors with similar K1 values. It is likely that this secretory granule α-amidation activity is involved in the physiological biosynthetic α-amidation of a wide range of bioactive peptides.  相似文献   

4.
3β-Hydroxysteroid dehydrogenase (3β-HSD)/Δ5→4-isomerase activity in steroidogenic tissues is required for the synthesis of biologically active steroids. Previously, by use of dehydroepiandrosterone (3β-hydroxy-5-androsten-17-one, DHEA) as substrate, it was established that in addition to steroidogenic tissues 3β-HSD/Δ5→4-isomerase activity also is expressed in extraglandular tissues of the human fetus. In the present study, we attempted to determine whether the C-5,C-6-double bond of DHEA serves to influence 3β-HSD activity. For this purpose, we compared the efficiencies of a 3β-hydroxy-5-ene steroid (DHEA) and a 3β-hydroxy-5α-reduced steroid (5α-androstane-3β,17β-diol, 5α-A-diol) as substrates for the enzyme. The apparent Michaelis constant (Km) for 5α-A-diol in midtrimester placenta, fetal liver, and fetal skin tissues was at least one order of magnitude higher than that for DHEA, viz the apparent Km of placental 3β-HSD for 5α-A-diol was in the range of 18 to 40 μmol/l (n = 3) vs 0.45 to 4 μmol/l for DHEA (n = 3); for the liver enzyme, 17 μmol/l for 5α-A-diol and 0.60 μmol/l for DHEA, and for the skin enzyme 14 and 0.18 μmol/l, respectively. Moreover, in 13 human fetal tissues evaluated the maximal velocities obtained with 5α-A-diol as substrate were higher than those obtained with DHEA. A similar finding in regard to Kms and rates of product formation was obtained by use of purified placental 3β-HSD with DHEA, pregnenolone, and 3β-hydroxy-5α-androstan-17-one (epiandrosterone) as substrates: the Km of 3β-HSD for DHEA was 2.8 μmol/l, for pregnenolone 1.9 μmol/l, and for epiandrosterone 25 μmol/l. The specific activity of the purified enzyme with pregnenolone as substrate was 27 nmol/mg protein·min and, with epiandrosterone, 127 nmol/mg protein·min. With placental homogenate as the source of 3β-HSD, DHEA at a constant level of 5 μmol/l behaved as a competitive inhibitor when the radiolabeled substrate, [3H]5α-A-diol, was present in concentrations of 20 to 60 μmol/l, but a lower substrate concentrations the inhibition was of the mixed type; similar results were obtained with [3H]DHEA as the substrate at variable concentrations in the presence of a fixed concentration of 5α-A-diol (40 μmol/l). These findings are indicative that both steroids bind to a common site on the enzyme, however, the binding affinity for these steroids appear to differ markedly as suggested by the respective Kms. Studies of inactivation of purified placental 3β-HSD/Δ5→4-isomerase by an irreversible inhibitor, viz 5,10-secoestr-4-yne-3,10,17-trione, were suggestive that the placental protein adopts different conformations depending on whether the steroidal substrate has a 5α-configuration, e.g. epiandrosterone, or a C-5,C-6-double bond e.g. DHEA or pregnenolone. The lower rates of product formation obtained with placenta and fetal tissues by use of 3β-hydroxy-5-ene steroids as substrates when compared with those obtained with 3β-hydroxy-5α-reduced steroids may be explained by a combination of factors, including: (i) inhibition of 3β-HSD activity by end products of metabolism of 3β-hydroxy-5-ene steroids, e.g. 4-androstene-3,17-dione formed with DHEA as substrate; (ii) higher binding affinity of the enzyme for 3β-hydroxy-5-ene steroids—and possibly for their 3-oxo-5-ene metabolites; (iii) lack of a requirement for the isomerization step with 5α-reduced steroids as substrates, and (iv) the possible presence in fetal tissues of an enzyme with 3β-HSD activity only (i.e. no Δ5→4-isomerase).  相似文献   

5.
α7β1 is the major integrin complex expressed in differentiated muscle cells where it functions as a laminin receptor. In this work we have expressed the α7 integrin subunit in CHO cells to investigate the functional properties of this receptor. After transfection with α7 CHO cells acquired the ability to adhere and spread on laminin 1 consistent with the laminin receptor activity of the α7β1. α7 transfectants, however, showed a 70% reduction in the ability to adhere to fibronectin and were unable to assemble a fibronectin matrix. The degree of reduction was inversely related to the level of α7 expression. To define the mechanisms underlying this adhesive defect we analyzed surface expression and functional properties of the α5β1 fibronectin receptor. Although cell surface expression of α5β1 was reduced by a factor of 20–25% in α7 transfectants compared to control untransfected cells, this slight reduction was not sufficient to explain the dramatic reduction in cell adhesion (70%) and matrix assembly (close to 100%). Binding studies showed that the affinity of125I-fibronectin for its surface receptor was decreased by 50% in α7 transfectants, indicating that the α5β1 integrin is partially inactivated in these cells. Inactivation can be reversed by Mn2+, a cation known to increase integrin affinity for their ligands. In fact, incubation of cells with Mn2+restored fibronectin binding affinity, adhesion to fibronectin, and assembly of fibronectin matrix in α7 transfectants. These data indicate that α7 expression leads to the functional down regulation of α5β1 integrin by decreasing ligand binding affinity and surface expression. In conclusion, the data reported establish the existence of anegative cooperativitybetween α7 and α5 integrins that may be important in determining functional regulation of integrins during myogenic differentiation.  相似文献   

6.
Type 3 17β-hydroxysteroid dehydrogenase (17β-HSD), a key steroidogenic enzyme, transforms 4-androstene-3,17-dione (Δ4-dione) into testosterone. In order to produce potential inhibitors, we performed solid-phase synthesis of model libraries of 3β-peptido-3α-hydroxy-5α-androstan-17-ones with 1, 2, or 3 levels of molecular diversity, obtaining good overall yields (23–58%) and a high average purity (86%, without any purification steps) using the Leznoff's acetal linker. The libraries were rapidly synthesized in a parallel format and the generated compounds were tested as inhibitors of type 3 17β-HSD. Potent inhibitors were identified from these model libraries, especially six members of the level 3 library having at least one phenyl group. One of them, the 3β-(N-heptanoyl- -phenylalanine- -leucine-aminomethyl)-3α-hydroxy-5α-androstan-17-one (42) inhibited the enzyme with an IC50 value of 227 nM, which is twice as potent as the natural substrate Δ4-dione when used itself as an inhibitor. Using the proliferation of androgen-sensitive (AR+) Shionogi cells as model of androgenicity, the compound 42 induced only a slight proliferation at 1 μM (less than previously reported type 3 17β-HSD inhibitors) and, interestingly, no proliferation at 0.1 μM.  相似文献   

7.
The immunologic cross-reactivity of the α and α+ forms of the large subunit and the β subunit of the (Na+ + K+)-ATPase from brain and kidney preparations was examined using rabbit antiserum prepared against the purified holo lamb kidney enzyme. As previously reported by Sweadner ((1979) J. Biol. Chem. 254, 6060–6067) phosphorylation of the large subunit of the (Na+ + K+)-ATPase in the presence of Na+, Mg2+, and [γ-32P]ATP revealed that dog and, very likely, rat brain contain two forms of the large subunit (designated α and α+) while dog, rat, and lamb kidney contain only one form (α). The cross-reactivity of the α and α+ forms in these preparations was investigated by resolving the subunits by SDS-polyacrylamide gel electrophoresis. The separated polypeptides were transferred to unmodified nitrocellulose paper, and reacted with rabbit anti-lamb kidney serum, followed by detection of the antigen-antibody complex with 125I-labeled protein A and autoradiography. By this method, the α and α+ forms of rat and dog brain, as well as the α form found in kidney, were shown to cross-react. In addition, membranes from human cerebral cortex were shown to contain two immunoreactive bands corresponding to the α and α+ forms of dog brain. In contrast, the brain of the insect Manduca sexta contains only one immunoreactive polypeptide with a molecular weight intermediate to the α and α+ forms of dog brain. The β subunit from lamb, dog and rat kidney and from dog and rat brain cross-reacts with anti-lamb kidney (Na+ + K+)-ATPase serum. The mobility of the β subunit from dog and rat brain on SDS-polyacrylamide electrophoresis gels is greater than the mobility of the β subunit from lamb, rat or dog kidney.  相似文献   

8.
In anterior pituitaries from male rats, it appeared that 5α-androstane-3β, 17β-diol was quickly metabolized into 5α-androstane-3β,6α-17β-triol and 5α-androstane-3β,7α, 17β-triol by action of 6α- and 7α-hydroxylases. Hydroxysteroid hydroxylases were located in endoplasmic reticulum and were dependent on NADPH+. Their optimum pH was 8.0, optima temperature, 37°C, and their apparent Km was 2.7 μM. Hydroxylative reactions were not reversible and not modified by gonadectomy. Hydroxylation seemed an efficient control of the pituitary level of 5α-andros-tane-3β, 17β-diol.  相似文献   

9.
Receptors for α2-macroglobulin-proteinase complexes have been characterized in rat and human liver membranes. The affinity for binding of 125I-labelled α2-macroglobulin · trypsin to rat liver membranes was markedly pH-dependent in the physiological range with maximum binding at pH 7.8–9.0. The half-time for association was about 5 min at 37°C in contrast to about 5 h at 4°C. The half-saturation constant was about 100 pM at 4°C and 1 nM at 37°C (pH 7.8). The binding capacity was approx. 300 pmol per g protein for rat liver membranes and about 100 pmol per g for human membranes. Radiation inactivation studies showed a target size of 466 ± 71 kDa (S.D., n = 7) for α2-macroglobulin · trypsin binding activity. Affinity cross-linking to rat and human membranes of 125I-labelled rat α1-inhibitor-3 · chymotrypsin, a 210 kDa analogue which binds to the α2-macroglobulin receptors in hepatocytes (Gliemann, J. and Sottrup-Jensen, L. (1987) FEBS Lett. 221, 55–60), followed by SDS-polyacrylamide gel electrophoresis, revealed radioactivity in a band not distinguishable from that of cross-linked α2-macroglobulin (720 kDa). This radioactivity was absent when membranes with bound 125I-α1-inhibitor-3 complex were treated with EDTA before cross-linking and when incubation and cross-linking were carried out in the presence of a saturating concentration of unlabelled complex. The saturable binding activity was maintained when membranes were solubilized in the detergent 3-[(3-cholamidopropyl)dimethylammonio]profane sulfonate (CHAPS) and the size of the receptor as estimated by cross-linking experiments was shown to be similar to that determined in the membranes. It is concluded that liver membranes contain high concentrations of an approx. 400–500 kDa α2-macroglobulin receptor soluble in CHAPS. The soluble preparation should provide a suitable material for purification and further characterization of the receptor.  相似文献   

10.
Laminins assemble into trimers composed of α, β, and γ chains which posttranslationally are glycosylated and sometimes proteolytically cleaved. In the current paper we set out to characterize posttranslational modifications and the laminin isoforms formed by laminin α1 and α5 chains. Comparative pulse–chase experiments and deglycosylation studies in JAR cells established that the Mr 360,000 laminin α1 chain is glycosylated into a mature Mr 400,000 band while the Mr 370,000 laminin α5 chain is glycosylated into a Mr 390,000 form that upon secretion is further processed into a Mr 380,000 form. Hence, despite the shorter peptide length of α1 chain in comparison with the α5 chain, secreted α1 assumes a larger size in SDS–PAGE due to a higher degree of N-linked glycosylation and due to the lack of proteolytic processing. Immunoprecipitations and Western blotting of JAR laminins identified laminin α1 and laminin α5 chains in laminin-1 and laminin-10. In placenta laminin α1 chain (Mr 400,000) and laminin α5 chain (Mr 380,000/370,000 doublet) were found in laminin-1/-3 and laminin-10/-11. Immunohistochemically we could establish that the laminin α1 chain in placenta is deposited in the developing villous and trophoblast basement membrane, also found to contain laminin β2 chains. Surprisingly, a fraction of the laminin α1 chain from JAR cells and placenta could not be precipitated by antibodies to laminin β1–β3 chains, possibly pointing to an unexpected complexity in the chain composition of α1-containing laminin isoforms.  相似文献   

11.
(Methyl 2-acetamidoacrylate)tricarbonyliron(0) (3) reacts with 2 equivalents of methyllithium to give methyl N-acetylalaninate (4) and 2-acetamido-4-oxopentanoate (5) when the reaction is quenched with trifluoroacetic acid. Production of methyl N-acetylalaninate is dependent only on the presence of trifluoroacetic acid, and the ratio of 4 to 5 generated in these reactions is related to the quantity of trifluoroacetic acid used to quench them. Addition of two equivalents of methyllithium followed by tertiary haloalkanes gives protected β,β,β-trialkyl α-amino acids which may be hydrolysed to give tert-leucine (13) and the new α-amino acids 2-amino-3,3-dimethylpentanoic acid (14) and 2-amino-3,3-dimethylhexanoic acid (15).  相似文献   

12.
A highly sensitive, kinetically unambiguous assay for α-factor-induced delay of cell passage through the “start” step of cell division in yeast is presented. The assay employs perfusion with periodic microscopy to monitor the bud emergence kinetics on the 20% of cells within an exponentially growing population which exist prior to the α-factor execution point of start. The t1/2 for cell passage through start by this population of cells is 31 min in the absence of α-factor. The inhibition constant, KI, represents the α-factor concentration which produces a 50% inhibition of this rate and is equal to 2×10−10M. A second assay for maximal cell division arrest by α-factor on whole populations of cells is presented. This assay shows a maximum cell division arrest time of 125±5 h at saturating α-factor, and a K50 (that is, an α-factor concentration which produces a half-maximal response) of 2.5×10−8M. Both assays were performed in the effective absence of α-factor inactivation. Values of the dissociation constant KD and total number of receptors per cell which specifically mediate cell division arrest or delay were estimated to be 2.5×10−8M and 104, respectively. These estimates, along with the quantitative dose-response data for division arrest which are presented here, are consistent with each receptor·α-factor complex which is present on the cell at equilibrium producing a 43±10 s delay of cell passage through start. Surprisingly, this number is constant within twofold over the entire range of cellular division arrest responses to α-factor, that is, from a 1.9-fold inhibition of the rate of cell passage through start at 0.17 nM α-factor to a 125±5 h maximum arrest at saturating α-factor concentrations of >170 nM. The possible significance of this observation toward the mechanism of α-factor-induced cell division arrest is discussed.  相似文献   

13.
An N-carbamoyl-β-alanine amidohydrolase of industrial interest from Agrobacterium tumefaciens C58 (βcarAt) has been characterized. βcarAt is most active at 30°C and pH 8.0 with N-carbamoyl-β-alanine as a substrate. The purified enzyme is completely inactivated by the metal-chelating agent 8-hydroxyquinoline-5-sulfonic acid (HQSA), and activity is restored by the addition of divalent metal ions, such as Mn2+, Ni2+, and Co2+. The native enzyme is a homodimer with a molecular mass of 90 kDa from pH 5.5 to 9.0. The enzyme has a broad substrate spectrum and hydrolyzes nonsubstituted N-carbamoyl-α-, -β-, -γ-, and -δ-amino acids, with the greatest catalytic efficiency for N-carbamoyl-β-alanine. βcarAt also recognizes substrate analogues substituted with sulfonic and phosphonic acid groups to produce the β-amino acids taurine and ciliatine, respectively. βcarAt is able to produce monosubstituted β2- and β3-amino acids, showing better catalytic efficiency (kcat/Km) for the production of the former. For both types of monosubstituted substrates, the enzyme hydrolyzes N-carbamoyl-β-amino acids with a short aliphatic side chain better than those with aromatic rings. These properties make βcarAt an outstanding candidate for application in the biotechnology industry.  相似文献   

14.
Structures of the complexes (η3-C3H5)Pd(μ-η6:1-CH2PhCr(CO)3 and (η3-C3H5)Pd[μ-η6:1-CH(Ph)Ph]Cr(CO)3 in solution were evaluated by NMR (1H and 13C) and IR spectroscopy. The dynamic behaviour of the complexes was investigated. Quick rotation (on the NMR time scale) of the tricarbonylchromium groups around the axis passing through the centre of the η6-coordinated phenyl ring and the chromium atom takes place at room temperature and becomes slow on cooling. The η3-allylic ligand was proved to undergo no dynamic changes in solution. Unlike the solid state, the semi-bridging carbonyl groups between chromium and palladium atoms are absent or very weak in solution. Cross-coupling reactions of the complexes with organohalides are described.  相似文献   

15.
This method involves the irreversible formation of a complex between 125I-labeled α-bungarotoxin and the acetylcholine receptor in either its membrane-bound or purified state. The separation of the labeled toxin-receptor complex from unreacted toxin is accomplished by chromatography on carboxymethylcellulose cation-exchange resin. The method described was developed to satisfy the following experimental requirements that could not be dealt with in their entirety by employing any of the published methods: (i) the complete recovery of reacted and unreacted species in relatively small volumes; (ii) an efficient and precise isolation of the specific and irreversible 125I-labeled α-bungarotoxin-receptor complex when the complexation reactions demand a large excess of unlabeled α-bungarotoxin for quenching (a 20-fold molar excess of unlabeled over labeled toxin); (iii) this isolation of the toxin-receptor complex allows one to determine the protein concentrations in the samples, a necessity in experiments covering a wide range of receptor concentrations; (iv) a consistent low blank for binding site concentrations ranging over two or three orders of magnitude; (v) simplicity and rapidity.  相似文献   

16.
The fate of 6α- and 6β-hydrogens of lathosterol during the transformation into 20-hydroxyecdysone was chased by feeding [3α,6β-2H2]- and [3α,6α-2H2]-lathosterols to hairy roots of Ajuga reptans var. atropurpurea. The behavior of 6β-hydrogen, which mostly migrated to the C-5 position of 20-hydroxyecdysone, was in agreement with that of C-6 hydrogen of cholesterol. The results strongly supported the view that cholesterol and lathosterol are first metabolized into 7-dehydrocholesterol, which is then converted into 20-hydroxyecdysone via 7-dehydrocholesterol 5α,6α-epoxide in the hairy roots.  相似文献   

17.
Radioimmunoassays for measuring prostaglandin F (PGF) and 5α, 7α-dihydroxy-11-keto tetranorprosta-1,16-dioic acid, PGF-main urinary metabolite (PGF-MUM), with 125I-tyrosine methylester amide (TMA) of PGF and PGF-MUM were developed.Antibody to PGF was produced in rabbits immunized with conjugates of PGF coupled to bovine serum albumine. Antibody to PGF-MUM was also produced in rabbits immunized with conjugates of PGF-MUM coupled to bovine serum albumin.PGF-125I-TMA had an affinity to antiserum to PGF. PGF-MUM-125I-TMA also responded to antiserum to PGF-MUM.  相似文献   

18.
From the cytosol fraction (supernatant fluid at 105,000 g) of chicken liver, 4-en-3-oxosteroid 5β-reductase (EC 1.3.1.23) was purified by ammonium sulfate precipitation, followed by Butyl Toyopearl, DEAE-Sepharose, Sephadex G-75 and hydroxylapatite column chromatographies. The enzyme activity was quantitated from amount of the 5β-reduced metabolites derived from [4-14C]testosterone. During the purification procedures, 17β-hydroxysteroid dehydrogenase which was present in the cytosol fraction was separated from 5β-reductase fraction by the Butyl Toyopearl column chromatography. By the DEAE-Sepharose column chromatography, 3α- and 3β-hydroxysteroid dehydrogenases were able to be removed from 5β-reductase fraction. The final enzyme preparation was apparently homogenous on SDS-polyacrylamide gel electrophoresis. Purification was about 13,600-fold from the hepatic cytosol. The molecular weight of this enzyme was estimated as 37,000 Da by SDS-polyacrylamide gel electrophoresis and also by Sephadex G-75 gel filtration. For 5β-reduction of 4-en-3-oxosteroids, such as testosterone, androstenedione and progesterone, NADPH was specifically required as cofactor. Km of 5β-reductase for NADPH was estimated as 4.22 × 10−6M and for testosterone, 4.60 × 10−6M. The optimum pH of this enzyme ranged from pH 5.0 to 6.5 and other enzymic properties of the 5β-reductase were examined.  相似文献   

19.
20.
Transforming growth factor type β (TGFβ) is a pleiotropic regulator of cell growth with specific high-affinity cell-surface receptors on a large number of cells; its mechanism of action, however, is poorly defined. In this report, we utilized the mouse fibroblast line AKR-2B to explore the question of the temporal requirements during the cell cycle in regard to both the growth inhibitory and the growth stimulatory action of TGFβ. The results indicate that AKR-2B cells are most sensitive to the inhibitory action of TGFβ during early to mid-G1. In addition, TGFβ need be present only briefly (as little as l min) in order to exert its inhibitory effect on EGF-induced DNA synthesis. Likewise, the stimulatory effect of TGFβ in the absence of EGF requires only an equally brief exposure to TGFβ. Use of homogeneous 125I-labeled TGFβ in a cell-binding assay demonstrates that TGFβ bound to cell-surface receptors can readily exchange into the culture medium T1/2 = 120 min), helping to rule out the possibility that persistent receptor-bound TGFβ is the source of a continuous stimulus. The data indicate that TGFβ exposure induces a stable state in the cell (T1/2 = 20 h) similar to but distinct from the state of “competence” induced by platelet-derived growth factor (PDGF).  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