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 Hussain  D Lewis  M Yu  P W Melera 《Gene》1992,112(2):179-188
Simian virus 40 promoter-enhancer-based mammalian expression plasmids using dihydrofolate reductase (DHFR)-encoding cDNA sequences originally isolated from two methotrexate (MTX)-resistant, DHFR-overproducing Chinese hamster lung cell lines were constructed. One, designated pSVA75, contains a DHFR cDNA that encodes leucine (Leu22) and corresponds to the wild type (wt), MTX-sensitive form of the enzyme [Melera et al., J. Biol. Chem. 263 (1988) 1978-1990]. The other plasmid, pSVA3, contains a cDNA that encodes a novel mutant form of the enzyme in which Leu22 has been changed to Phe [Melera et al., Mol. Cell Biol. 4 (1984) 38-48]. The resulting DHFR displays a 20-fold-enhanced resistance to inhibition by MTX, but maintains the catalytic activity of the wt enzyme [Albrecht et al., Cancer Res. 32 (1972) 1539-1546]. Transfection of DHFR- Chinese hamster ovary cells with either plasmid demonstrated that both were able to reconstitute the DHFR+ phenotype with equal efficiency (i.e., greater than 2.5 x 10(-3), indicating that both the wt and mutant enzymes were catalytically active in transfected cells. In addition, the mutant form of the enzyme was found to act as a dominant selectable marker when transfected into diploid DHFR+ cells, and to allow selection of resistant clones at low MTX concentrations (125 nM MTX) with a frequency of greater than 8 x 10(-4). Moreover, transfected clones were found to amplify their exogenous DHFR sequences to reasonably high levels (42-fold) at relatively low (888 nM) MTX concentrations, suggesting that substantial amplification of DHFR DNA and cotransfected sequences as well, can be achieved with this vector.  相似文献   

2.
3.
The enzyme, dihydrofolate reductase (DHFR), from Mycobacterium tuberculosis (mt-DHFR) is believed to be a potential drug target for the treatment of tuberculosis. The co-crystal structure of mt-DHFR bound with glycerol (GOL), NAPDH and methotrexate (MTX) reveals a GOL binding site on the enzyme. This GOL binding site could be very important for the design of novel, selective mt-DHFR inhibitors, because this binding site is absent on human DHFR (h-DHFR). We have performed molecular dynamic simulations and free energy calculations to evaluate the binding affinity of GOL and its free energy contribution to the binding of MTX to mt-DHFR. The results showed that GOL does not bind tightly to mt-DHFR. Although GOL itself contributed free energy on MTX binding to mt-DHFR, GOL also increased the flexibilities of MTX, so that MTX cannot maintain strong electronic interactions with ARG32 and ARG60, which caused the total binding free energy to decrease. These data suggest that GOL binding is weak and it could be expelled from the binding site, to allow inhibitors containing appropriate side chains to bind. This observation can be used to inform future drug design studies, especially those aimed at improving drug selectivity against h-DHFR.  相似文献   

4.
5.
Summary A petunia cell line, 1ECB, was previously isolated by the stepwise selection procedure, for resistance to methotrexate (MTX), an antimetabolite for the enzyme dihydrofolate reductase (DHFR). Using ammonium sulfate precipitates of cell lysates of cell line 1ECB and its parental cell line (WT), it was found that the mutant has an increase of 400 fold in 3H-MTX binding capacity and a decrease in the affinity for MTX binding, at two orders of magnitude, in comparison with the WT. In addition, the DHFR specific activity in the mutant increased only moderately (5- to 10-fold), this activity is extremely sensitive to MTX inhibition as compared to the WT. It is evident that the MTX resistance of line 1ECB results mainly from overproduction of an MTX-binding protein which differs from the WT DHFR by four biochemical criteria. This protein may serve as a trap for the excess amounts of MTX to which the cells are exposed.  相似文献   

6.
Polyclonal antibodies against dihydrofolate reductase (DHFR) from the human lymphoblastoid cell line WIL-2/M4 were used as probes to compare the antigenic structures in solution of native DHFRs obtained from a broad range of species and their complexes with substrate, cofactor, and folate antagonist inhibitors. All these antibodies could bind to the denatured human DHFR, indicating that they were specific for the primary structure of this enzyme. Denatured chicken liver and L1210 murine leukemic DHFRs competed for all of the antibodies that bound to the human enzyme, although less effectively than the denatured human enzyme, showing the presence of similar epitopes among the vertebrate enzymes. However, both direct binding and competition experiments showed low antibody cross-reactivities with native chicken liver (8%) and murine (10%) DHFRs, suggesting differences in the disposition of similar epitopes in these enzymes. The lactobacillus casei DHFR showed a low amount (less than 2%) of cross-reactivity with the antibodies while the same antibodies did not cross-react with the Escherichia coli enzyme. DHFR from soybean seedlings competed for a large proportion (70%) of the anti-human DHFR antibodies, indicating a close similarity in the antigenic structures of plant and animal DHFRs. Binary complexes of the L. casei, avian, murine, and human DHFRs with dihydrofolate, methotrexate (MTX), trimethoprim (TMP), NADPH, and NADP+ all showed significantly lower antibody binding capacity as compared with the corresponding free enzymes. Further, these ligands inhibited antibody binding to the enzyme to varying degrees.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
8.
Dihydrofolate reductase (DHFR; EC 1.5.1.3) was purified to homogeneity from soybean seedlings by affinity chromatography on methotrexate-aminohexyl Sepharose, gel filtration on Ultrogel AcA-54, and Blue Sepharose chromatography. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the enzyme gave a single protein band corresponding to a molecular weight of 22,000. The enzyme is not a 140,000 Da heteropolymer as reported by others. Amino acid sequence-specific antibodies to intact human DHFR and also antibodies to CNBr-generated fragments of human DHFR bound to the plant enzyme on Western blots and cross-reacted significantly in immunoassays, indicating the presence of sequence homology between the two enzymes. The plant and human enzymes migrated similarly on nondenaturing polyacrylamide electrophoretic gels as monitored by activity staining with a tetrazolium dye. The specific activity of the plant enzyme was 15 units/mg protein, with a pH optimum of 7.4. Km values of the enzyme for dihydrofolate and NADPH were 17 and 30 microM, respectively. Unlike other eukaryotic enzymes, the plant enzyme showed no activation with organic mercurials and was inhibited by urea and KCl. The affinity of the enzyme for folate was relatively low (I50 = 130 microM) while methotrexate bound very tightly (KD less than 10(-10) M). Binding of pyrimethamine to the plant enzyme was weaker, while trimethoprim binding was stronger than to vertebrate DHFR. Trimetrexate, a very potent inhibitor of the human and bacterial enzymes showed weak binding to the plant enzyme. However, certain 2,4-diaminoquinazoline derivatives were very potent inhibitors of the plant DHFR. Thus, the plant DHFR, while showing similarity to the vertebrate and bacterial enzymes in terms of molecular weight and immunological cross-reactivity, can be distinguished from them by its kinetic properties and interaction with organic mercurials, urea, KCl and several antifolates.  相似文献   

9.
Interaction of the pore-forming protein (porin) from Yersinia pseudotuberculosis with S- and R-forms of the endogenous lipopolysaccharide (LPS) was studied at various ionic strengths (20-600 mM NaCl), concentrations of divalent cations (5-100 mM CaCl2, MgCl2), and pH values from 3.0 to 9.0. The interaction of the R-LPS with porin has been shown in all experimental conditions to be in consensus with the model suggesting binding at independent sites of two types. S-LPS binds to interacting sites of relatively high affinity and to independent sites of low affinity at all pH values examined and at low NaCl concentration. The cooperative interaction of the S-LPS and porin is not observed at high ionic strength and in divalent cation-free medium. The number of binding sites of porin and association constants (Ka) for both LPS forms decrease significantly on increasing the solution ionic strength. The Ka values for the R- and S-LPS change oppositely on changing the pH: the Ka value for the R-LPS is maximal (Ka = 6.7 x 10(5) M-1), but that for S-LPS is minimal (Ka = 0.4 x 10(5) M(-1) at pH 5.0-5.5. The number of high-affinity and low-affinity binding sites for both LPS forms is maximal at pH 5.0-5.5. In this case, the numbers of high- and low-affinity sites for R-LPS are 3 and 10, respectively, and those for the S-LPS are 7 and 20, respectively. These data suggest an important role of electrostatic interactions on binding of LPS to porin. The contribution of conformational changes of the ligand and protein and hydrophobic interactions are discussed.  相似文献   

10.
The self-association in aqueous solution of folic acid (FA), 7,8-dihydrofolic acid (DHFA) and 5,6,7,8-tetrahydrofolic acid (THFA) has been studied by the use of proton magnetic resonance (1H NMR) spectroscopy. At concentrations below 10 mM, all three folates exist in (monomer)2 in equilibrium dimer equilibria with association constants (Ka) equal to 400, 66 and 14 M-1 for FA, DHFA and THFA respectively. These values decreased markedly to 157, 18 and 3 M-1, for FA, DHFA and THFA respectively, in the presence of 0.8 M KCl. The high extent of dimerization of FA is believed to impede the interaction with the active site of dihydrofolate reductase (DHFR) rendering it a poor substrate. In contrast, the DHFA with a much lower Ka is a better substrate. Conditions that lower the Ka of both FA and DHFA, (i.e., 0.8M KCl) turn them into better substrates. Based on the findings of the present study, it is also predicted that dihydro MTX may be a better inhibitor of DHFR than MTX.  相似文献   

11.
R L Blakley  L Cocco 《Biochemistry》1985,24(17):4704-4709
Stopped-flow spectrophotometry and stopped-flow fluorometry have been used to study the binding of methotrexate (MTX) and 3-deazamethotrexate (3-deazaMTX) to dihydrofolate reductase (DHFR) isoenzymes from Streptococcus faecium and from Lactobacillus casei. The absorbance change and fluorescence quenching that occur when MTX binds to DHFR isoenzyme II from S. faecium (SFDHFR II) are both biphasic and give similar apparent rate constants for both phases. The faster phase has an apparent rate constant that is dependent on MTX concentration and therefore corresponds to the initial binding reaction. From the concentration dependence it has been calculated that the association rate constant is 3.0 X 10(5) M-1 s-1 at 20 degrees C and pH 7.3, and the association constant (equilibrium constant) under these conditions is 5.8 X 10(5) M-1. By examination of the amplitude of the fast-phase absorbance change at various wavelengths, it has been determined that the absorbance change occurring in the fast phase is due to MTX protonation. Within the limits of the method it was thus not possible to detect a difference in the rates of binding and of protonation of MTX. The MTX association rate constant is pH dependent, decreasing 330-fold as the pH is decreased from 5.0 to 9.0. The data fit well to a curve generated by assuming a single ionization with a pKa of 6.0 and a pH-independent association rate constant 1000-fold greater for binding of protonated MTX to SFDHFR II than for binding of unprotonated MTX.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Dihydrofolate reductase (DHFR) and thymidylate synthase (TS) activities are associated with a 285,000 molecular weight enzyme complex in carrot (Daucus carota L.). Selection for methotrexate (MTX) resistance by stepwise increase of the concentration of MTX results in a high frequency adaptation to MTX with little or no significant increase in DHFR activity. However, when as a second step following MTX selection a specific inhibitor of TS, 5-fluoro-2-deoxyuridine was used, DHFR overproducer lines were obtained. The overproduction phenotype of the lines was almost completely lost after 8 weeks of growth in the absence of selection pressure. Although DHFR and TS are independent gene products, their activities increase in proportion (~20-fold) in the overproducer lines. This strongly suggests that DHFR and TS are not only functionally and physically linked in the same enzyme complex, but also are coregulated. These cell lines resemble the MTX-induced DHFR overproducer amplified cell lines of mammalian origin in their mode of selection, high frequency of appearance, elevated enzyme activity, and increased specific mRNA levels.  相似文献   

13.
The potential for ligand-initiated signal transduction through B cell membrane IgM is assessed in terms of ligand concentration, binding site valency, and binding site affinity for membrane Ig. Estimates of the physicochemical requirements for achieving G0* enhancement of class II MHC expression, G1 entry, and S phase entry in human B cells were made by comparing the stimulatory effects of three affinity-diverse anti-Cmu2 mAb when in bivalent (unconjugated) form, or as mAb-dextran conjugates with low binding site valency (oligovalent ligands) or high binding site valency (multivalent ligands). An increase in binding site number (and concomitant molecular mass) caused a profound reduction in both the minimal concentration and affinity requisites for B cell activation. The enhancing effect of increased binding site valency was most evident for the signaling of those most distal stages in B cell activation, i.e., G1 and S phase, which were difficult to induce with bivalent ligands. The results suggest that highly multimeric TI-2 Ag may be good immunogens because they are able to elicit a full activation response not only from infrequent high affinity B cells, but also from a substantial proportion of the many lower affinity Ag-specific B cells in virgin B cell populations. Interestingly, the activation of B cells by ligands with binding sites of high intrinsic affinity (Ka = 5 x 10(8) M-1) was less influenced by increases in binding site valency than was B cell activation by ligands with intermediate binding site affinity (Ka = 2 x 10(7) M-1). This suggests that the minimal epitope valency requirement for T cell-independent B cell activation by mIg cross-linking Ag may be dependent on the intrinsic affinity with which membrane Ig molecules on a given B cell interact with the redundantly expressed epitopes.  相似文献   

14.
The binding of N3- to Co(II)-substituted bovine carbonic anhydrase III was measured at various pH values by spectrophotometric titrations. The apparent Ki values were found to increase with pH in the studied range between pH 5.8 and 8.9. The inhibition of CO2 hydration by N-3 was found to be essentially uncompetitive at all investigated pH values (pH 6.3-8.9). The Ki values for the inhibition of kcat are much smaller than those obtained in the spectrophotometric titrations indicating that an enzyme form with a high affinity for N-3, presumably having a metal-bound H2O, accumulates in the steady state at saturating CO2 concentrations. Assuming that the low pH limit of Ki = 9 microM for the inhibition of kcat represents the affinity of N-3 for the Co(II)-OH2 form, a pKa value near 5 can be estimated for Co(II)-bound water from the pH dependence of N-3 binding in the absence of CO2. Measurements of time-resolved absorption spectra during CO2 hydration in the presence of a low N-3 concentration showed the transient appearance of the characteristic spectrum of the enzyme-N-3 adduct clearly demonstrating the accumulation in the steady state of an enzyme form with a high affinity for N-3. In similar experiments without inhibitor the transient formation of a spectral form corresponding to a Co(II)-OH2 species has been demonstrated. This spectral form is rather featureless lacking the absorption maxima at 618 nm and 640 nm characteristic of the Co(II)-OH- species. Our results strongly support the hypothesis that the rate-limiting step in CO2 hydration catalyzed by carbonic anhydrase III is the protolysis of metal-bound water.  相似文献   

15.
We developed a method to determine dihydrofolate reductase (DHFR) activity at pH 7.4 (37 degrees C) by monitoring its product, tetrahydrofolate (H(4)folate), using HPLC with electrochemical detection. After the assay mixture was deproteinized by 0.5 M perchloric acid, the H(4)folate concentration was measured. Using sodium ascorbate at 20 mM, H(4)folate was stable in our assay system. The enzyme activity was also stable. The detection limit of this method was less than 1 nM of H(4)folate in the enzyme assay system, which was 1/100 lower than those for the NADPH-spectrophotometric assay, which is commonly used for analysis of DHFR activity. This value of 1 nM allowed us to control the conversion from dihydrofolate (H(2)folate) to H(4)folate less than 10% of initial substrate concentrations during assay, when we used a concentration around K(m) values reported for DHFR from various sources. The rate of reduction showed a linearity at concentrations around the K(m). The reduction rate must be evaluated exactly around the K(m), in order to obtain an accurate profile of Michaelis-Menten kinetics. This assay method has a sensitivity high enough to determine the reduction rate at H(2)folate concentrations around K(m). In addition, the assay procedure is very simple. Therefore, our method may be useful for studying DHFR.  相似文献   

16.
Folate cycle kinetics in human breast cancer cells   总被引:1,自引:0,他引:1  
  相似文献   

17.
Synthesis and biological evaluation of a fluorescent analogue of folic acid   总被引:4,自引:0,他引:4  
A fluorescein derivative of the lysine analogue of folic acid, N alpha-pteroyl-N epilson-(4'-fluoresceinthiocarbamoyl)-L-lysine (PLF), was synthesized as a probe for dihydrofolate reductase (DHFR) and a membrane folate binding protein (m-FBP). Excitation of PLF at 282 nm and at 497 nm gave a fluorescence emission maximum at 518 nm. Binding of PLF to human DHFR or human placental m-FBP results in approximately a 20-fold enhancement in the magnitude of the fluorescence emission, suggesting that the ligand interacts with a hydrophobic region on these proteins. Additional evidence suggests that an energy transfer may occur between the pteridine and the fluorescein moieties. PLF binds to the active site of human DHFR since methotrexate (MTX) competes stoichiometrically and the denatured enzyme in the presence of PLF did not exhibit fluorescent enhancement. The dissociation constant for the fluorescein derivative with respect to human DHFR is 115 nM as compared to 111 nM for folic acid. The Ki value for the competitive inhibition of human DHFR by the fluorescent analogue of folic acid is 2.0 microM compared to 0.48 microM for folic acid. PLF was reduced to N alpha-(7,8-dihydropteroyl)-N epilson-(4'-fluoresceinthiocarbamoyl)-L-lysine (H2PLF) and assayed by the enzymatic conversion to the tetrahydro derivative. The Km value for human DHFR for the dihydrofolate analogue is 2.0 microM. The KD value for H2PLF to human DHFR is 47 nM as compared to 44 nM for dihydrofolate. The KD values for both H2PLF and PLF indicate that the fluorescein moiety does not significantly affect folate binding in enzyme binary complexes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
To investigate the relationship between thin filament Ca2+ binding and activation of the MgATPase rate of myosin subfragment 1, native cardiac thin filaments were isolated and characterized. Direct measurements of 45Ca binding to the thin filament were consistent with non-cooperative binding to two high affinity sites (Ka 7.3 +/- 0.8 x 10(6) M-1) and either cooperative or non-cooperative binding to one low affinity site (Ka 4 +/- 2 x 10(5) M-1) per troponin at 25 degrees C, 30 mM ionic strength, pH 7.06. Addition of a low concentration of myosin subfragment 1 to the native thin filaments produced a Ca2+-regulated MgATPase activity with Kapp (2.5 +/- 1.3 x 10(5) M-1), matching the low affinity Ca2+ site. The MgATPase rate was cooperatively activated by Ca2+ (Hill coefficient 1.8). To determine whether Ca2+ binding to the low affinity sites was cooperative, native thin filament troponin was exchanged with troponin labeled on troponin C with 2-(4'-iodoacetamidanilo)naphthalene-6-sulfonic acid. From the Ca2+-sensitive fluorescence of this complex, Ca2+ binding was cooperative with a Hill coefficient of 1.7-2.0. Using the troponin-exchanged thin filaments, myosin subfragment 1 MgATPase rate activation was also cooperative and closely proportional to Ca2+ thin filament binding. Reconstitution of the thin filament from its components raised the Ca2+ affinity by a factor of 2 (compared with native thin filaments) and incorporation of fluorescently modified troponin raised the Ca2+ affinity by another factor of 2. Stoichiometrically reconstituted thin filaments produced non-cooperative MgATPase rate activation, contrasting with cooperative activation with native thin filaments, troponin-exchanged thin filaments and thin filaments reconstituted with a stoichiometric excess of troponin. The Ca2+-induced fluorescence transition of stoichiometrically reconstituted thin filaments was non-cooperative. These results suggest that Ca2+ binds cooperatively to the regulatory sites of the cardiac thin filament, even in the absence of myosin, and even though cardiac troponin C has only one Ca2+-specific binding site. A theoretical model for these observations is described and related to the experimental data. Well-known interactions between neighboring troponin-tropomyosin complexes are the proposed source of cooperativity and also influence the overall Ka. The data indicate that Ca2+ is four times more likely to elongate a sequence of troponin-tropomyosin units already binding Ca2+ than to bind to a site interior to a sequence of units without Ca2+.  相似文献   

19.
  • 1.1. Covalent coupling of fluorescein to methotrexate (MTX) by a 5-carbon spacer yields a dihydrofolate reductase (DHFR) inhibitor (FMTX) with Ki = 11 nM.
  • 2.2. FMTX shows a fluorescence quenching with respect to fluorescein which is relieved by binding to the enzyme.
  • 3.3. The dissociation constants (Kd) of MTX, FMTX, NADPH and 7,8-dihydrofolate (DHF) from bovine liver DHFR have been determined by fluorometric titrations.
  • 4.4. The Kd values for NADPH, MTX and FMTX from the complementary binary complexes (MTX·DHFR, FMTX·DHFR and NADPH·DHFR) were also obtained; these show a 2- to 4-fold decrease with respect to those obtained by titration of the free enzyme.
  • 5.5. A competitive assay for MTX has been developed by exploiting the fluorescence enhancement of DHFR-bound FMTX. This assay may be useful for the routine determination of MTX in the concentration range from 10−9 to 10−7 M.
  相似文献   

20.
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