首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The partial amino acid sequence of dihydrofolate reductase (DHFR, EC 1.5.1.3) from human KB/6b cells has been determined by using 3.5 mg of protein. Peptides covering the entire polypeptide chain were recovered from preparative peptide maps generated by the combination of paper chromatography and electrophoresis at pH 4.4 Peptide maps from mouse L1210 DHFR were also generated for comparison. Amino acid sequence of 75% of the 186 amino acid residues in the polypeptide chain of human KB/6b DHFR was obtained from Edman degradations and the remaining sequence was deduced from the amino acid compositions, from electrophoretic mobilities of related peptides and from the sequence homologies with other known mammalian DHFR sequences. A comparison of the proposed human DHFR sequence with the previously known sequences of mouse enzyme [Stone, et al. (1979) J. Biol. Chem. 245, 480-488] indicates that 18 differences are located in the established sequence of 139 residues and that 5 additional differences are in the tentative sequence of the remaining 47 amino acids. Kinetic properties of human KB/6b and mouse L1210 DHFR, which were determined in parallel experiments, are also compared. The possible structural-functional relationships between human and mouse DHFR are discussed.  相似文献   

2.
Cody V  Pace J  Chisum K  Rosowsky A 《Proteins》2006,65(4):959-969
Structural data are reported for two highly potent antifolates, 2,4-diamino-5-[3',4'-dimethoxy-5'-(5-carboxy-1-pentynyl)]benzylpyrimidine (PY1011), with 5000-fold selectivity for Pneumocystis carinii dihydrofolate reductase (pcDHFR), relative to rat liver DHFR, and 2,4-diamino-5-[2-methoxy-5-(4-carboxybutyloxy)benzyl]pyrimidine (PY957), that has 80-fold selectivity for pcDHFR. Crystal structures are reported for NADPH ternary complexes with PY957 and pcDHFR, refined to 2.2 A resolution; with PY1011 and pcDHFR, refined to 2.0 A resolution; and with PY1011 and mouse DHFR (mDHFR), refined to 2.2 A resolution. These results reveal that the carboxylate of the omega-carboxyalkyloxy side chain of these inhibitors form ionic interactions with the conserved Arg in the substrate binding pocket of DHFR. These data suggest that the enhanced inhibitory activity of PY1011 compared with PY957 is, in part, due to the favorable contacts with Phe69 of pcDHFR by the methylene carbons of the inhibitor side chain that are oriented by the triple bond of the 1-pentynyl side chain. These contacts are not present in the PY957 pcDHFR complex, or in the PY1011 mDHFR complex. In the structure of mDHFR the site of Phe69 in pcDHFR is occupied by Asn64. These data also revealed a preference for an unusual parallel ring stacking interaction between Tyr35 of the active site helix and Phe199 of the C-terminal beta sheet in pcDHFR and by Tyr33 and Phe179 in mDHFR that is independent of bound ligand. A unique His174-His187 parallel ring stacking interaction was also observed only in the structure of pcDHFR. These ring stacking interactions are rarely found in any other protein families and may serve to enhance protein stability.  相似文献   

3.
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   

4.
The transient state kinetics of catalysis for dihydrofolate reductase (DHFR) from several enzyme sources including highly purified recombinant human enzyme (rHDHFR) have been examined. Like DHFR from Escherichia coli, the enzyme from Lactobacillus casei, and isoenzyme 2 from Streptococcus faecium exhibit a slow increase in activity upon addition of substrates to enzyme. No slow hysteresis of this type was detected with recombinant human DHFR (rHDHFR) or DHFR from chicken or bovine liver or L1210 mouse leukemia cells (MDHFR). In contrast, both rHDHFR and MDHFR exhibited a very rapid decrease in activity (t1/2 = 30 and 20 ms, respectively) during a phase that occurred after the first turnover of the enzyme but before establishment of the steady state. This intermediate phase was not observed for the bacterial enzymes or the avian enzyme, nor was it observed with a mutant of rHDHFR in which Phe-31 has been replaced by leucine. For rHDHFR the intermediate phase is not a consequence of product inhibition, substrate depletion, or enzyme instability. It may therefore be concluded that this unusual transient state kinetic behavior results from the existence of two conformers of the enzyme, one of which has a higher turnover number than the other with the equilibrium shifting in favor of the less active conformer during the course of catalysis. The equilibrium is particularly favorable for the less active conformer when NADP is present in the active site of rHDHFR, whereas bound tetrahydrofolate favors the more active conformer. The more active conformer has a 6-fold higher Km for dihydrofolate than does the less active conformer. The existence of these conformers is likely to produce cooperative behavior by rHDHFR in vivo.  相似文献   

5.
The x-ray structure of the PTX:NADPH:L22F human mutant DHFR ternary complex was used as a structural template to generate structural models for the following wild type DHFR complexes: PTX:DHFR:NADPH, TMP:DHFR:NADPH, EPM:DHFR:NADPH, and TMQ:DHFR:NADPH. Each of these complexes were subsequently modeled in a 60 Å cube of explicit water and minimized to a rms gradient of from 1.0-3.0·10-5 kcal·Å-1. For each complex, interaction energies were calculated for the antifolate interaction with each of the following: the DHFR binding site residues, the entire DHFR protein, the solvated complex (containing DHFR, NADPH, and solvent water), water alone, and NADPH. Additionally, each antifolate was subdivided into distinct substructural regions and interaction energy calculations were performed in order to evaluate their contributions to overall antifolate interaction. Each antifolate showed its most stable interaction with the solvated complex. Substructural regions which consisted of a nitrogen containing aromatic ring system contributed most to the stability of the antifolate interactions, while the hydrocarbon aromatic rings, methoxy, and ethoxy groups showed much less stable interaction energies. Since the different substructural regions of nonclassical antifolates differ in their contributions to overall antifolate binding, those substructural regions which exhibit relatively unfavorable interaction energies may constitute important targets in the design of improved DHFR inhibitors.  相似文献   

6.
When NADPH was added in excess to a bovine liver DHFR solution, a fluorescence peak due to an energy transfer mechanism was apparent at 450 nm. It did not vary over time. The intrinsic fluorescence peak of DHFR at 320 nm was quenched and this phenomenon increased over the time-course after NADPH addition. This result was ascribed to a slow DHFR conformational change induced by NADPH binding, which has never been previously described in such a long time scale (more than 30 min). A kinetic scheme accounting for this mechanism has been proposed. Furthermore, this interconversion between two protein conformers led to an increase in the initial apparent rate of the enzymatic reaction catalyzed by DHFR.  相似文献   

7.
One of the most widely used antitumor drugs is cis-diamminedichloroplatinum(II) (cisplatin), and mechanisms of cisplatin resistance have been investigated in numerous model systems. Many studies have used mouse leukemia L1210/0 as a reference wild-type cell line, and cisplatin-resistant subclones have been derived from it. Increased DNA excision repair capacity is thought to play a key role in the acquired cisplatin resistance, and this has influenced development of drugs for clinical trials. We report here that the L1210/0 line is in fact severely deficient in nucleotide excision repair of damaged DNA in vivo and in vitro. L1210/0 cell extracts could be complemented by extracts from repair-defective human xeroderma pigmentosum (XP) or rodent excision repair cross-complementing (ERCC) mutant cells, except for XPG/ERCC5 mutants. Purified XPG protein could restore repair proficiency to L1210/0 extracts. Expression of mouse XPG mRNA was similar in all L1210 lines studied, suggesting a point mutation or small alteration of XPG in L1210/0 cells. The DNA repair capacity of a cisplatin-resistant subline, L1210/DDP10, is similar to that of type culture collection L1210 cells and to those of other normal mammalian cell lines. Nucleotide excision repair of DNA is thus clearly important in the intrinsic cellular defense against cisplatin. However, in contrast to what is generally believed, enhancement of DNA repair above the normal level in these rodent cells does not appear to be a mechanism of acquired resistance to the drug.  相似文献   

8.
The crystal structure of calmodulin (Mr 16,700, 148 residues) from Drosophila melanogaster as expressed in a bacterial system has been determined and refined at 2.2-A resolution. Starting with the structure of mammalian calmodulin, we produced an extensively refitted and refined model with a conventional crystallographic R value of 0.197 for the 5,239 reflections (F greater than or equal to 2 sigma (F)) within the 10.0-2.2-A resolution range. The model includes 1,164 protein atoms, 4 calcium ions, and 78 water molecules and has root mean square deviations from standard values of 0.018 A for bond lengths and 0.043 A for angle distances. The overall structure is similar to mammalian calmodulin, with a seven-turn central helix connecting the two calcium-binding domains. The "dumb-bell" shaped molecule contains seven alpha-helices and four "EF hand" calcium-binding sites. Although the amino acid sequences of mammalian and Drosophila calmodulins differ by only three conservative amino acid changes, the refined model reveals a number of significant differences between the two structures. Superimposition of the structures yields a root mean square deviation of 1.22 A for the 1,120 equivalent atoms. The calcium-binding domains have a root mean square deviation of 0.85 A for the 353 equivalent atoms. There are also differences in the amino terminus, the bend of the central alpha-helix, and the orientations of some of the side chains.  相似文献   

9.
Human dihydrofolate reductase (DHFR) protein levels rapidly increase upon exposure to methotrexate, a potent inhibitor of this enzyme. A model to explain this increase proposes that DHFR inhibits its own translation by binding to its cognate mRNA and that methotrexate disrupts the DHFR protein-mRNA complex allowing its translation to resume. In the present study, Chinese hamster ovary cells lacking DHFR were transfected with wild type and mutants of human DHFR to identify amino acids that are essential for increases in DHFR in response to methotrexate. Glu-30, Leu-22, and Ser-118 were involved in the up-regulation of DHFR protein levels by methotrexate and certain other antifolates. Cells transfected with E30A, L22R, and S118A mutants that did not respond to methotrexate up-regulation had higher basal levels of DHFR, consistent with the model, i.e. lack of feedback regulation of these enzymes. Although cells containing the S118A mutant enzyme had higher levels of DHFR and had catalytic activity similar to that of wild type DHFR, they had the same sensitivity to the cytotoxicity of methotrexate, as were cells with wild type DHFR. This finding provides evidence that the adaptive up-regulation of DHFR by methotrexate contributes to the decreased sensitivity to this drug. Based on these observations, a new model is proposed whereby DHFR exists in two conformations, one bound to DHFR mRNA and the other bound to NADPH. The mutants that are not up-regulated by methotrexate are unable to bind their cognate mRNA.  相似文献   

10.
Isothermal titration calorimetry (ITC) was used to determine the thermodynamic driving force for inhibitor binding to the enzyme dihydrofolate reductase (DHFR) from Escherichia coli. 1,4-Bis-{[N-(1-imino-1-guanidino-methyl)]sulfanylmethyl}-3,6-dimethyl-benzene (1) binds DHFR:NADPH with a K(d) of 13±5 nM while the related inhibitor 1-{[N-(1-imino-guanidino-methyl)]sulfanylmethyl}-3-trifluoromethyl-benzene (2) binds DHFR:NADPH with a K(d) of 3.2±2.2 μM. The binding of these inhibitors has both a favorable entropy and enthalpy of binding. Additionally, we observe positive binding cooperativity between both 1 and 2 and the cofactor NADPH. Binding of compound 1 to DHFR is 285-fold tighter in the presence of the NADPH cofactor than in its absence. We did not detect binding of 2 to DHFR in the absence of NADPH. The backbone amide (1)H and (15)N NMR resonances of DHFR:NADPH and both DHFR:NADPH inhibitor complexes were assigned in order to better understand the binding of these inhibitors in solution. The chemical shift perturbations observed with the binding of 1 were greatest at residues closest to the binding site, but significant perturbations also occur away from the inhibitor location at amino acids in the vicinity of residue 58 and in the GH loop. The pattern of chemical shift changes observed with the binding of 2 is similar to that seen with 1. The main differences in chemical shift perturbation between the two inhibitors are in the Met20 loop and in residues at the interface between the inhibitor and NADPH.  相似文献   

11.
The binding of trimethoprim to dihydrofolate reductase from L1210 mouse lymphoma cells has been studied by measuring the changes in chemical shift of nuclei of the ligand that accompanying binding. The 6- and 2',6'-proton chemical shifts of bound trimethoprim have been determined by transfer of saturation experiments, and the 2-carbon chemical shift has been determined by using [2-13C]trimethoprim. The changes in proton chemical shift are substantially smaller than those accompanying binding to bacterial dihydrofolate reductase [Cayley, P. J., Albrand, J. P., Feeney, J., Robert, G. C. K., Piper, E. A., & Burgen, A. S. V. (1979) Biochemistry 18, 3886]. It is shown that this difference arises largely from the fact that trimethoprim adopts different conformations when bound to mammalian and to bacterial dihydrofolate reductase. The proton chemical shifts are interpreted in terms of ring-current contributions from the two aromatic rings of trimethoprim itself and the nearby aromatic amino acid residues of the enzyme. The latter have been located by using the refined crystallographic coordinates of the Lactobacillus casei and Escherichia coli reductases in their complexes with methotrexate [Bolin, J. T., Filman, D. J., Matthews, D. A. & Kraut, J. (1982) J. Biol. Chem. 257, 13650], under the assumption that, as indicated by the 13C chemical shifts, the diaminopyrimidine ring of trimethoprim binds in the same way as does the corresponding part of methotrexate. With use of these assumptions, the conformation of trimethoprim bound to the dihydrofolate reductases from L. casei, E. coli, and L1210 cells has been calculated.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Four spin-labeled inhibitors of dihydrofolate reductase (DHFR) have been synthesized, each of which has the 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) reporting group at a different distance from the 2,4-diaminopyrimidine moiety by which the inhibitors are anchored and oriented in the active site. Inhibitors in which the TEMPO group is attached by a short side chain are weakly bound to DHFR from bacteria (Streptococcus faecium and Lactobacillus casei), to the bovine enzyme and to recombinant human DHFR. However, binding is sufficiently tight, especially in the ternary complexes with NADPH, for recording of the EPR spectra of the bound ligands. The spectra indicate that when these inhibitors are bound to the enzyme the TEMPO group is highly immobilized with correlation time, tau c, 4-20ns. Inhibitors that have the reporter group attached to the glutamate moiety of methotrexate bind to all four DHFRs more tightly than the inhibitors with shorter side chains by factors of up to 10(6). However, in most complexes formed by the inhibitors with longer side chains immobilization of the TEMPO group is slight (tau c 0.2-4 ns). These results are in general agreement with predictions from X-ray crystallographic results including thermal factors but there are some unanticipated differences between some results for bacterial and eukaryotic enzymes. Three of the splin-labeled inhibitors would provide good probes for distance measurements in and around the active site of mammalian DHFR.  相似文献   

13.
Using cloned DNA complementary to mouse dihydrofolate reductase (DHFR) mRNA, the organization of the hamster DHFR gene has been determined in two baby hamster kidney (BHK) cell lines, A5 and B1. A5 cells are highly methotrexate-resistant, containing 200-fold more copies of the DHFR gene than do the parental B1 cells. The DHFR gene has the same organization in A5 and B1 cells, suggesting that it has not been altered by the amplification process. The BHK DHFR gene spans a maximum of 10.7 kb and contains at least three introns. Thus the BHK DHFR gene is much smaller than the mouse DHFR gene, which has a minimum size of 42 kb and at least five introns. This striking size difference is probably due to much smaller introns in the BHK DHFR gene.  相似文献   

14.
15.
Oligonucleotide-directed, site-specific mutagenesis was used to convert phenylalanine-31 of human recombinant dihydrofolate reductase (DHFR) to leucine. This substitution was of interest in view of earlier chemical modification studies (Kumar et al., 1981) and structural studies based on X-ray crystallographic data (Matthews et al., 1985a,b) which had implicated the corresponding residue in chicken liver DHFR, Tyr-31, in the binding of dihydrofolate. Furthermore, this particular substitution allowed testing of the significance of protein sequence differences between mammalian and bacterial reductases at this position with regard to the species selectivity of trimethoprim. Both wild-type (WT) and mutant (F31L) enzymes were expressed and purified by using a heterologous expression system previously described (Prendergast et al., 1988). Values of the inhibition constants (Ki values) for trimethoprim were 1.00 and 1.08 microM for WT and F31L, respectively. Thus, the presence of phenylalanine at position 31 in human dihydrofolate reductase does not contribute to the species selectivity of trimethoprim. The Km values for nicotinamide adenine dinucleotide phosphate (reduced) (NADPH) and dihydrofolate were elevated 10.8-fold and 9.4-fold, respectively, for the mutant enzyme, whereas the Vmax increased only 1.8-fold. Equilibrium dissociation constants (KD values) were obtained for the binding of NADPH and dihydrofolate in binary complexes with each enzyme. The KD for NADPH is similar in both WT and F31L, whereas the KD for dihydrofolate is 43-fold lower in F31L. Values for dihydrofolate association rate constants (kon) with enzyme and enzyme-NADPH complexes were measured by stopped-flow techniques.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The development of drug resistance is reducing the efficiency of antifolates as antimalarials. This phenomenon has been linked to the occurrence of mutations in the parasite's dihydrofolate reductase (DHFR). In this way, the resistance to pyrimethamine and cycloguanil, two potent inhibitors of P. falciparum DHFR, is mainly related to mutations (single and crossed) at residues 16, 51, 59, 108 and 164 of the enzyme. In this work, we have refined a recently proposed homology-model of P. falciparum DHFR, and the resulting structure was used to obtain models for 14 mutant enzymes, employing molecular modeling. Ternary complexes of the mutant enzymes with these inhibitors have been superimposed to equivalent ternary complexes of the wild-type enzyme, allowing the proposition of hypotheses for the role of each mutation in drug resistance. Based on these results, possible reasons for antifolate resistance have been proposed.  相似文献   

17.
We developed a strategy for finding out the adapted variants of enzymes, and we applied it to an enzyme, dihydrofolate reductase (DHFR), in terms of its catalytic activity so that we successfully obtained several hyperactive cysteine- and methionine-free variants of DHFR in which all five methionyl and two cysteinyl residues were replaced by other amino acid residues. Among them, a variant (M1A/M16N/M20L/M42Y/C85A/M92F/C152S), named as ANLYF, has an approximately seven times higher k(cat) value than wild type DHFR. Enzyme kinetics and crystal structures of the variant were investigated for elucidating the mechanism of the hyperactivity. Steady-state and transient binding kinetics of the variant indicated that the kinetic scheme of the catalytic cycle of ANLYF was essentially the same as that of wild type, showing that the hyperactivity was brought about by an increase of the dissociation rate constants of tetrahydrofolate from the enzyme-NADPH-tetrahydrofolate ternary complex. The crystal structure of the variant, solved and refined to an R factor of 0.205 at 1.9-angstroms resolution, indicated that an increased structural flexibility of the variant and an increased size of the N-(p-aminobenzoyl)-L-glutamate binding cleft induced the increase of the dissociation constant. This was consistent with a large compressibility (volume fluctuation) of the variant. A comparison of folding kinetics between wild type and the variant showed that the folding of these two enzymes was similar to each other, suggesting that the activity enhancement of the enzyme can be attained without drastic changes of the folding mechanism.  相似文献   

18.
The X-ray structures of red yeast Sporobolomyces salmonicolor carbonyl reductase (SSCR) and its complex with a coenzyme, NADPH, have been determined at a resolution of 1.8A and 1.6A, respectively. SSCR was crystallized in an orthorhombic system with the space group P2(1)2(1)2(1) and cell dimensions of a=54.86 A, b=83.49 A, and c=148.72 A. On its cocrystallization with NADPH, isomorphous crystals of the SSCR/NADPH complex were obtained. The structure of SSCR was solved by a single wavelength anomalous diffraction measurement using a selenomethionine-substituted enzyme, and that of the SSCR/NADPH complex was solved by a molecular replacement method using the solved structure of SSCR. The structures of SSCR and the SSCR/NADPH complex were refined to an R-factor of 0.193 (R(free)=0.233) and 0.211 (R(free)=0.238), respectively. SSCR has two domains, an NADPH-binding domain and a substrate-binding domain, and belongs to the short-chain dehydrogenases/reductases family. The structure of the NADPH-binding domain and the interaction between the enzyme and NADPH are very similar to those found in other structure-solved enzymes belonging to the short-chain dehydrogenases/reductases family, while the structure of the substrate-binding domain is unique. SSCR has stereoselectivity in its catalytic reaction, giving rise to excessive production of (S)-alcohols from ethyl 4-chloro-3-oxobutanoate. The X-ray structure of the SSCR/NADPH complex and preliminary modeling show that the formation of the hydrophobic channel induced by the binding of NADPH is closely related to the stereoselective reduction by SSCR.  相似文献   

19.
Nucleoside permeation across mammalian cell membranes is complex with at least four distinct transporters known. Two of these (es and ei) are equilibrative (facilitated diffusion) carriers that have been studied is considerable detail. The other two (cif and cit) are concentrative, Na(+)-dependent carriers. A major obstacle to the characterization of the latter two mechanisms has been the lack of suitable model systems expressing only a single nucleoside transport activity. The present study describes the isolation of a cell line that has cif as its sole nucleoside transporter. L1210/MC5-1 cells, which have es and cif transport activity, were mutagenized and plated in soft agar containing two cytotoxic nucleosides (tubercidin (7-deazaadenosine) and cytosine arabinoside) that are substrates for es but not cif. A clonal line (L1210/MA-27.1) was isolated which retained the capacity for Na(+)-dependent [3H]formycin B transport but was unable to transport [3H]thymidine, a substrate for es but not cif. Failure of the mutant to transport thymidine was also demonstrated by the inability of thymidine (with adenine as a purine source) to rescue these cells from methotrexate toxicity. Furthermore, the mutant lacked nitrobenzylthioinosine (NBMPR) binding activity (an integral part of the es transporter) as demonstrated by reversible NBMPR binding and photoaffinity labeling with [3H]NBMPR. Loss of es transport activity was also demonstrated by the failure of NBMPR to affect the toxicity of 2-chlorodeoxyadenosine (IC50 approximately 30 nM) in L1210/MA27.1 cells. In contrast, NBMPR decreased the IC50 for 2-chlorodeoxyadenosine from 100 to 30 nM in the parental L1210/MC5-1 cell line. These results are consistent with the mechanism of NBMPR potentiation of 2-chlorodeoxyadenosine toxicity in L1210 cells being a blockade of efflux via es while the nucleoside is pumped into the cells by the concentrative cif carrier.  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