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 0 毫秒
1.
A purification procedure is reported for obtaining bovine liver dihydrofolate reductase in high yield and amounts of 100-200 mg. A key step in the procedure is the use of an affinity gel prepared by coupling pteroyl-L-lysine to Sepharose. The purified reductase has a specific activity of about 100 units/mg and is homogeneous as judged by analytical ultracentrifugation, polyacrylamide gel electrophoresis, and titration with methotrexate. The products of the first step of Edman degradation indicated a minimum purity of 79%. The reductase has a molecular weight of about 21500 on the basis of amino acid composition and 22100 +/- 300 from equilibrium sedimentation. It is not inhibited by antiserum to the Streptococcus faecium reductase (isoenzyme 2). Unlike the reductase of many other vertebrate tissues, the bovine enzyme is inhibited by mercurials rather than activated and it has a single pH optimum at both low and high ionic strength. However, the position of the pH optimum is shifted and the activity increased by increasing ionic strength. Automatic Edman degradation has been used to determine 34 of the amino-terminal 37 amino acid residues. Considerable homology exists between this region and the corresponding regions of the reductase from S. faecium and from Escherichia coli. This strengthens the idea that this region contributes to the structure of the binding site for dihydrofolate.  相似文献   

2.
Dihydrofolate reductase (DHFR) has been purified over 30,000-fold from Drosophila adults with a yield of 35%, using a combination of low pH extraction, (NH4)2SO4 precipitation, Sephadex gel filtration, Affi-Gel blue affinity chromatography, ion exchange and gel filtration FPLC. The Drosophila enzyme is a soluble, 17-22 kDa monomeric protein displaying the two pH optima characteristic of eukaryotic DHFRs. The sequence of the first 23 amino acids from the amino-terminal end of the protein shows that Drosophila DHFR is more homologous to the mosquito and vertebrate DHFRs than to the prokaryotic enzymes. However, the percent similarity between the two insect enzymes is not as close as expected when compared to the virtually identical initial sequence conservation of mammalian DHFRs.  相似文献   

3.
4.
We constructed mouse dihydrofolate reductase (DHFR) minigenes (dhfr) that had 1.5 kilobases of 5' flanking sequences and contained either none or only one of the intervening sequences that are normally present in the coding region. They were greater than or equal to 3.2 kilobase long, about one-tenth the size of the corresponding chromosomal gene. Both of these minigenes complemented the DHFR deficiency in Chinese hamster ovary dhfr-1-cells at a high frequency after DNA-mediated gene transfer. The level of DHFR enzyme in various transfected clones varied over a 10-fold range but never was as high as in wild-type Chinese hamster ovary cells. In addition, the level of DHFR in primary transfectants did not vary directly with the copy number of the minigene, which ranged from fewer than five to several hundred per genome. The minigenes could be amplified to a level of over 2,000 copies per genome upon selection in methotrexate, a specific inhibitor of DHFR. In one case, the amplified minigenes were present in a tandem array; in two other cases, a rearranged minigene plasmid and its flanking chromosomal DNA sequence were amplified. Thus, the mouse dhfr minigenes could be transcribed, expressed, and amplified in Chinese hamster ovary cells, although the efficiency of expression was generally low. The key step in the construction of these minigenes was the generation in vivo of lambda phage recombinants by overlapping regions of homology between genomic and cDNA clones. The techniques used here for dhfr should be generally applicable to any gene, however large, and could be used to generate novel genes from members of multigene families.  相似文献   

5.
6.
A pentapeptide which potently inhibits primary IgE antibody formation, Asp-Ser-Asp-Gly-Lys (DSDGK), has been efficiently produced with the aid of the dihydrofolate reductase (DHFR) handle [M. Iwakura, et al. (1992) J. Biochem. 111, 37-45]. The genes coding fused proteins comprising DHFR and multimeric forms of DSDGK, namely, DHFR-(DSDGK)3, DHFR-(DSDGK)14, and DHFR-(DSDGK)28, were constructed and expressed in Escherichia coli. The C-terminal peptides attached to DHFR did not affect the expression or the function of the DHFR handle, even when the length of the C-terminal peptide was as long as 160 amino acid residues. The fused proteins were easily purified by methotrexate affinity chromatography, one of the major advantages of the DHFR handle. The fused proteins were digested with trypsin and the monomeric peptide, DSDGK, was purified by HPLC. The yields of the peptide were estimated to be 11, 43, and 99 mg per 1 gram of the total cell proteins from E. coli cells producing DHFR-(DSDGK)3, DHFR-(DSDGK)14, and DHFR-(DSDGK)28, respectively.  相似文献   

7.
One of the earliest steps in translation initiation is recognition of the mRNA cap structure (m7GpppX) by the initiation factor eIF4E. Studies of interactions between purified eIF4E and its binding partners provide important information for understanding mechanisms underlying translational control in normal and cancer cells. Numerous impediments of the available methods used for eIF4E purification led us to develop a novel methodology for obtaining fractions of eIF4E free from undesired by-products. Herein we report methods for bacterial expression of eIF4E tagged with mutant dihydrofolate reductase (DHFR) followed by isolation and purification of the DHFR–eIF4E protein by using affinity and anion exchange chromatography. Fluorescence quenching experiments indicated the cap-analog, 7MeGTP, bound to DHFR–eIF4E and eIF4E with a dissociation constant (Kd) of 6 ± 5 and 10 ± 3 nM, respectively. Recombinant eIF4E and DHFR–eIF4E were both shown to significantly enhance in vitro translation in dose dependent manner by 75% at 0.5 μM. Nevertheless increased concentrations of eIF4E and DHFR–eIF4E significantly inhibited translation in a dose dependent manner by a maximum at 2 μM of 60% and 90%, respectively. Thus, we have demonstrated that we have developed an expression system for fully functional recombinant eIF4E. We have also shown that the fusion protein DHFR–eIF4E is functional and thus may be useful for cell based affinity tag studies with fluorescently labeled trimethoprim analogs.  相似文献   

8.
9.
Dihydrofolate reductase has been purified from a methotrexate-resistant strain of Lactobacillus casei NCB 6375. By careful attention to growth conditions, up to 2.5 g of enzyme is obtained from a 400 litre culture. The purification procedure, involving poly-ethyleneimine treatment, DEAE-cellulose chromatography and affinity chromatography on methotrexate-aminohexyl-Sepharose, operates on the gram scale, with overall yields of 50-60%. Elution of the affinity column by reverse (upward) flow was used, as it led to recovery of the enzyme in a much smaller volume. The enzyme obtained appears to be more than 98% pure, as judged by gel electrophoresis, isoelectric focusing, and gel filtration. It has a mol.wt. of approx. 17900 and a turnover number of 4s-1 (50mM-triethanolamine/400mM-KCl, pH 7.2, 25 degrees C) with dihydrofolate and NADPH as substrates. The turnover number for folate is 0.02s-1. Michaelis constants for a variety of substrates have been measured by using a new fluorimetric assay (0.36 muM-dihydrofolate; 0.78 muM-NADPH), and binding constants determined by using the quenching of protein fluorescence (dihydrofolate, 2.25 X 10(6)M-1; NADPH, greater than 10(8)M-1). The pH/activity profile shows a single maximum at pH 7.3; at this pH, marked activation by 0.5M-NaCl is observed.  相似文献   

10.
Chicken liver dihydrofolate reductase purified to apparent homogeneity by affinity chromatography contains tightly bound dihydrofolate. The most effective method for removal of the bound substrate is by electrofocusing. This procedure also removes previously unsuspected contaminants. In addition, the isoelectric profile revealed as many as four distinct peaks of enzyme activity. The major peak (pI = 8.4) represents 60–75% of the total activity, is devoid of bound substrate, and exhibits an A280A260 ratio approaching 1.9 and a specific activity of 14 units/mg. The peak of activity at the isoelectric point of 7.4 contains bound dihydrofolate. The major isoelectric band is shown to be homogeneous by the usual criteria. Notable features of the amino acid composition include a single cysteine, three tryptophans, and an excess of acidic residues. The N-terminal residue is valine. The molecular weight as determined by sedimentation equilibrium is 22,474. The s20,w0 is 2.07. A frictional coefficient of 1.2 indicates that the enzyme approximates a sphere. Circular dichroism measurements suggest a low α-helical content and a high degree of β-structure. The molar extinction coefficient was determined to be 28,970.  相似文献   

11.
The R-factor R388 mediates the production of a trimethoprim-resistant dihydrofolate reductase. This enzyme has a different molecular weight and pH profile to the trimethoprim-sensitive enzyme of the Escherichia coli host. The R-factor mediated enzyme was separated completely from the host E. coli enzyme by DEAE-cellulose ion-exchange chromatography. The purified R-factor enzyme was about 20 000 times less susceptible to trimethoprim than the E. coli enzyme and although it was inhibited competitively by trimethoprim, its inhibitor constant (Ki) was 20 000 times greater than that of the host enzyme. The R388 and E. coli enzymes also differed in their substrate specificity requirements. In addition, the R388 enzyme suprisingly conferred high level resistance to the broad spectrum dihydrofolate reductase inhibitor, amethopterin. The possible origins of the R388 enzyme are discussed.  相似文献   

12.
R plasmid dihydrofolate reductase with subunit structure.   总被引:3,自引:0,他引:3  
Dihydrofolate reductase, specified by the type II plasmid of a trimethoprim-resistant Escherichia coli, was purified 40-fold to homogeneity using a combination of gel filtration, DEAE-Sephacel chromatography, and hydrophobic chromatography. The final product shows a single protein band on polyacrylamide gel electrophoresis and has a specific activity of 1.0 unit/mg. The molecular weight of the purified enzyme is 36,000 as determined both by gel filtration and Ferguson analysis of polyacrylamide gel electrophoresis. In contrast, a single polypeptide with a molecular weight of 8,500 was observed on sodium dodecyl sulfate-gel electrophoresis. These experiments suggest that, unlike any bacteria or vertebrate dihydrofolate reductase previously examined, the type II R plasmid reductase is a tetramer composed of four identical subunits. A partial amino acid sequence determination shows no heterogeneity of the subunits and also no clear homology with any reductase sequence previously reported.  相似文献   

13.
14.
利用PCR技术从大肠杆菌DH5α中获取二氢叶酸还原酶(DHFR)基因folA。用限制性内切酶BamHI与PstI将该片段插入到克隆载体pUC18上,DNA测序鉴定目的基因。而后再将该基因亚克隆到表达载体pTrcHisC上,IPTG诱导表达重组蛋白。在非变性条件下,用TALON金属亲和层析树脂纯化含组氨酸标记的重组DHFR。纯化产物在热诱导条件下行SDSPAGE分析,除23000大小的单体外,还出现了交联的二聚体和多聚体;而当反应体系中含有还原剂β-巯基乙醇时,二聚体和多聚体都被减弱。推断蛋白质在热诱导条件下二级结构发生改变而产生交联,并且有二硫键的参与。  相似文献   

15.
The biological potency of the synthetic replicates of three peptides isolated from a human pancreatic tumor with growth hormone releasing activity and rat hypothalamic growth hormone-releasing factor was evaluated in conscious freely-moving rats and anesthetized rats. All 4 peptides are equipotent on a molar basis in their ability to stimulate GH secretion. Studies with synthetic fragments of the human derived material indicated that the amino-terminal amino acid is required for activity. Deletion of as many as 13 amino acids from the carboxy-terminal failed to decrease GH-releasing activity; however, deletion of 16 amino acids resulted in a significant decrease and deletion of 20 amino acids resulted in complete loss of bioactivity.  相似文献   

16.
R67 is a Type II dihydrofolate reductase (DHFR) that catalyzes the reduction of dihydrofolate (DHF) to tetrahydrofolate by facilitating the addition of a proton to N5 of DHF and the transfer of a hydride ion from NADPH to C6. Because this enzyme is a plasmid-encoded DHFR from trimethoprim-resistant bacteria, extensive studies on R67 with various methods have been performed to elucidate its reaction mechanism. Here, Raman difference measurements, conducted on the ternary complex of R67.NADP(+).DHF believed to be an accurate mimic of the productive DHFR.NADPH.DHF complex, show that the pK(a) of N5 in the complex is less than 4. This is in clear contrast to the behavior observed in Escherichia coli DHFR, a substantially more efficient enzyme, where the pK(a) of bound DHF at N5 is increased to 6.5 compared with its solution value of 2.6. A comparison of the ternary complexes in R67 and E. coli DHFRs suggests that enzymic raising of the pK(a) at N5 can significantly increase the catalytic efficiency of the hydride transfer step. However, R67 shows that even without such a strategy an effective DHFR can still be designed.  相似文献   

17.
18.
The binding of substrates and inhibitors to dihydrofolate reductase was studied by steady-state kinetics and high-field 1H-n.m.r. spectroscopy. A series of 5-substituted 2,4-diaminopyrimidines were examined and were found to be 'tightly binding' inhibitors of the enzyme (Ki less than 10(-9) M). Studies on the binding of 4-substituted benzenesulphonamides and benzenesulphonic acids also established the existence of a 'sulphonamide-binding site' on the enzyme. Subsequent n.m.r. experiments showed that there are two binding sites for the sulphonamides on the enzyme, one of which overlaps the coenzyme (NADPH) adenine-ring-binding site. An examination of the pH-dependence of the binding of sulphonamides to the enzyme indicated the influence of an ionizable group on the enzyme that was not directly involved in the sulphonamide binding. The change in pKa value from 6.7 to 7.2 observed on sulphonamide binding suggests the involvement of a histidine residue, which could be histidine-28.  相似文献   

19.
Primary cell cultures were prepared from fetal, neonatal and adult rat pituitaries and evaluated for their ability to secrete growth hormone (GH) in response to growth hormone-releasing factor (GRF). Pituitary cells prepared from fetuses at days 19 and 21 of gestation, neonatal animals at the day of birth (day 0) or the following day (day 1) and peripubertal male rats showed full dose response curves to GRF with maximal GH release when stimulated with 1 X 10(-10) M rat GRF. At this concentration of GRF, the amount of GH released was not different from that elicited by activation of adenylate cyclase with 1 X 10(-5) M forskolin. In contradistinction, a preparation of cells from fetuses at day 18 of gestation did not show the same release of GH when challenged with 1 X 10(-10) M GRF and forskolin (0.057 +/- 0.001, compared to 0.076 +/- 0.003 micrograms/10(5) cells per 4.5 h), although the cells clearly responded to both secretagogues (basal levels of GH, 0.029 +/- 0.002 micrograms/10(5) cells per 4.5 h). While cells prepared from fetuses at day 21 of gestation or from animals after birth released 5-10% of their total cellular GH content, those prepared from 18- and 19-day fetuses released as much as 40% of their total GH suggesting there is a maturation of intracellular GH processing that occurs late in gestation. The results show that, in late pregnancy, the rat fetal pituitary is highly responsive to growth hormone-releasing factor and suggest that this peptide participates in regulating GH levels during the perinatal period.  相似文献   

20.
A mouse complementary deoxyribonucleic acid segment coding for the enzyme dihydrofolate reductase has been cloned in two general classes of vectors containing simian virus 40 deoxyribonucleic acid: (i) those that can be propagated as virions in permissive cells and (ii) those that can be introduced into and maintained stably in various mammalian cells. Both types of vectors express the mouse dihydrofolate reductase by using signals supplied by simian virus 40 deoxyribonucleic acid sequences. Moreover, plasmid vectors carrying the complementary deoxyribonucleic acid segment can complement Chinese hamster ovary cells lacking dihydrofolate reductase.  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