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 46 毫秒
1.
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   

2.
Peptides from human dihydrofolate reductase (DHFR) generated by cyanogen bromide cleavage and corresponding to residues 15-52, 53-111, 112-125, and 140-186 (carboxyl terminus) were purified and used to immunize rats. Titration of the immune sera against denatured human DHFR by solid-phase immunoassay showed that peptides 15-52 and 140-186 were relatively highly immunogenic, unlike the native enzyme which is most immunogenic in the sequence 53-111. The antisera were specific for the corresponding peptides used for immunization. Antibodies to peptides 15-52, 53-111, and 140-186 cross-reacted with native human DHFR in solution in competition assays. However, the binding of nicotinamide adenine dinucleotide phosphate (reduced) (NADPH) and the inhibitors folate and methotrexate, both in binary and in ternary complexes with the enzyme, caused a striking reduction in binding of antibody. Using a sensitive radioactive assay, it was found that antisera to peptides 15-52 and 140-186, both of which exhibited a high antibody titer, caused significant inhibition of DHFR. Because peptide 140-186 does not include any active-site residues, it is concluded that at least in this case all the antibodies bound to regions outside the active site. Since comparison of the X-ray structures of the chicken liver DHFR holoenzyme with the apoenzyme reveals no changes in secondary structural elements (alpha-helices and beta-sheets), the reduction in antibody binding to DHFR-ligand complexes must not involve epitopes within these structures.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Substrate and inhibitor binding to dihydrofolate reductase (DHFR) primarily involves residues in the amino-terminal half of the enzyme; however, antibody binding studies performed in this laboratory suggested that the loop region located in the carboxyl terminus of human DHFR (hDHFR; residues 140-186) is involved in conformational changes that occur upon ligand binding and affect enzyme function (Ratnam, M., Tan, X., Prendergast, N.J., Smith, P.L. & Freisheim, J.H. (1988) Biochemistry 27, 4800-4804). To investigate this observation further, site-directed mutagenesis was used to construct deletion mutants of hDHFR missing 1 (del-1), 2 (del-2), 4 (del-4), and 6 (del-6) residues from loops in the carboxyl terminus of the enzyme. The del-1 mutant enzyme has a two-amino acid substitution in addition to the one-amino acid deletion. Deletion of only one amino acid resulted in a 35% decrease in the specific activity of the enzyme. The del-6 mutant enzyme was inactive. Surprisingly, the del-4 mutant enzyme retained a specific activity almost 33% that of the wild type. The specific activity of the del-2 mutant enzyme was slightly higher (38% wild-type activity) than that of the del-4 mutant. All three active deletion mutants were much less stable than the wild-type enzyme, and all three showed at least a 10-fold increase in Km values for both substrates. The del-1 and del-2 mutants exhibited a similar increase in KD values for both substrate and cofactor. The three active deletion mutants lost activity at concentrations of activating agents such as KCl, urea, and p-hydroxymercuribenzoate that continued to stimulate the wild-type enzyme. Antibody binding studies revealed conformational differences between the wild-type and mutant enzymes both in the absence and presence of bound folate. Thus, although the loops near the carboxyl terminus are far removed from the active site, small deletions of this region significantly affect DHFR function, indicating that the loop structure in mammalian DHFR plays an important functional role in its conformation and catalysis.  相似文献   

4.
Human transforming growth factor alpha (TGF alpha) is a 50-residue mitogenic peptide with a compact structure restrained by three disulfide bonds. Sequential and overlapping synthetic peptides were made to identify epitopes of TGF alpha using a panel of murine monoclonal antibodies and rabbit polyclonal antibodies. Antibodies were raised against human TGF alpha from different preparations obtained from either chemical synthesis or recombinant DNA techniques. Two related methodologies were used in these experiments. In the first method, probes were synthesized as peptides immobilized on polyethylene pins by the method of Geysen et al. (Geysen, H. M., Meloen, R. H., and Barteling, S. J. (1984) Proc. Natl. Acad. Sci. U.S.A. 81, 3998-4002). Three sets of sequentially overlapping tetrapeptides, hexapeptides, and octapeptides covering the entire length of the human TGF alpha sequence were synthesized. In the second method, a set of overlapping 8-residue synthetic peptides, freely soluble in solution, were used as probes. By both methods, the nonneutralizing monoclonal antibodies, i.e. those that did not inhibit TGF alpha in mitogenic assays, recognized two immunodominant regions represented by the NH2-terminal segment (residues 1-9) and the most prominent beta-sheet of the molecule (residues 22-31). The NH2 terminus and the beta-sheet-(22-31) are in the same face of the molecule as determined by the solution structure. These two immunodominant regions were also recognized by the polyclonal antibodies as well as regions in the COOH terminus as minor epitopes. However, none of the neutralizing monoclonal antibodies recognized any synthetic peptides. Thus, our results suggest that the receptor-binding surface of TGF alpha does not involve the face represented by the NH2-terminal fragment and the major beta-sheet of residues 22-31, but rather, that the opposite face represented by two loops formed by residues 12-20 and 34-43 may be involved in TGF alpha binding to its receptor.  相似文献   

5.
6.
Polyclonal and monoclonal antibodies were generated against a synthetic peptide (25 amino acid residues) corresponding to the amino acid sequence surrounding the active site serine of Torpedo californica acetylcholinesterase (AChE). Prior to immunization, the peptide was either coupled to bovine serum albumin or encapsulated into liposomes containing lipid A as an adjuvant. To determine whether this region of AChE is located on the surface of the enzyme and thus accessible for binding to antibodies, or located in a pocket and thus not accessible to antibodies, the immunoreactivity of the antibodies was determined using enzyme-linked immunosorbent assay (ELISA), immunoprecipitation, Western blots, and competition ELISA. The polyclonal antibody and several of the monoclonal antibodies failed to react with either Torpedo or fetal bovine serum AChE in their native conformations, but showed significant cross-reactivity with the denatured enzymes. Human serum butyrylcholinesterase, which has a high degree of amino acid sequence homology with these AChEs, failed to react with the same antibodies in either native form or denatured form. Chymotrypsin also failed to react with the monoclonal antibodies in either form. Eighteen octapeptides spanning the entire sequence of this region were synthesized on polyethylene pins, and epitopes of representative monoclonal antibodies were determined by ELISA. The reactivity of peptides suggest that a portion of the 25 mer peptide in AChE containing the active site serine is the primary epitope. It is not exposed on the surface of the enzyme and is most likely sequestered in a pocket-like conformation in the native enzyme.  相似文献   

7.
8.
There is marked pH dependence of the rate constant (koff) for tetrahydrofolate (H4folate) dissociation from its ternary complex with human dihydrofolate reductase (hDHFR) and NADPH. Similar pH dependence of H4folate dissociation from the ternary complex of a variant of hDHFR with the substitution Phe31----Leu (F31L hDHFR) causes this dissociation to become rate limiting in the enzyme mechanism at pH approximately 5, and this accounts for the marked decrease in kcat for this variant as the pH is decreased from 7 to 5. This decreased kcat at low pH is not seen for most DHFRs. koff for dissociation of folate, dihydrofolate (H2folate), and H4folate from their binary complexes with hDHFR is similarly pH dependent. For all the complexes examined, the pH dependence of koff in the range pH 5-7 is well described by a pKa of about 6.2 and must be due to ionization of a group on the enzyme. In the higher pH range (7-10), koff increases further as the pH is raised, and this relation is governed by a second pKa which is close to the pKa for ionization of the amide group (HN3-C4O) of the respective ligands. Thus, ionization of the ligand amide group also increases koff. Evidence is presented that the dependence of pH on koff for hDHFR accounts for the shape of the kcat versus pH curve for both hDHFR as well as its F31L variant and contributes to the higher efficiency of hDHFR compared with bacterial DHFR.  相似文献   

9.
Growth of methotrexate-resistant dihydrofolate reductase gene-amplified KB cells in the presence of 5-fluorouracil results in an increase in dihydrofolate reductase mRNA. This increase can be solely attributed to a species of RNA of approximately 3.5 kilobase pairs in size. Although dihydrofolate reductase enzyme activity increases per cell with increasing 5-fluorouracil, there is a decrease of enzyme activity per mg of protein (Dolnick, B. J., and Pink, J. J. (1983) J. Biol. Chem. 258, 13299-13306). The rate of in vivo enzyme synthesis, as assayed by immunoprecipitation and supported by gel electrophoresis, does not decrease and may in fact increase with increasing 5-fluorouracil. Translation of purified dihydrofolate reductase mRNA in vitro shows that the rate of translation is unaffected by 5-fluorouracil incorporation into mRNA. The inhibition of dihydrofolate reductase by a monospecific polyclonal antiserum is reduced with extracts from 5-fluorouracil-treated cells. Inhibition of dihydrofolate reductase by methotrexate is significantly reduced in extracts from 5-fluorouracil-treated cells compared to control extracts. Tight binding of [3H]methotrexate is also different in extracts from 5-fluorouracil-treated cells. This data supports the hypothesis of translational miscoding during protein synthesis as a major mechanism of 5-fluorouracil-mediated cytotoxicity and suggests a new mechanism of 5-fluorouracil-methotrexate antagonism.  相似文献   

10.
The crystal structures of two human dihydrofolate reductase (hDHFR) ternary complexes, each with bound NADPH cofactor and a lipophilic antifolate inhibitor, have been determined at atomic resolution. The potent inhibitors 6-([5-quinolylamino]methyl)-2,4-diamino-5-methylpyrido[2,3-d]pyrimidine (SRI-9439) and (Z)-6-(2-[2,5-dimethoxyphenyl]ethen-1-yl)-2,4-diamino-5-methylpyrido[2,3-d]pyrimidine (SRI-9662) were developed at Southern Research Institute against Toxoplasma gondii DHFR-thymidylate synthase. The 5-deazapteridine ring of each inhibitor adopts an unusual puckered conformation that enables the formation of identical contacts in the active site. Conversely, the quinoline and dimethoxybenzene moieties exhibit distinct binding characteristics that account for the differences in inhibitory activity. In both structures, a salt-bridge is formed between Arg70 in the active site and Glu44 from a symmetry-related molecule in the crystal lattice that mimics the binding of methotrexate to DHFR.  相似文献   

11.
Escherichia coli dihydrofolate reductase (DHFR) has several flexible loops surrounding the active site that play a functional role in substrate and cofactor binding and in catalysis. We have used heteronuclear NMR methods to probe the loop conformations in solution in complexes of DHFR formed during the catalytic cycle. To facilitate the NMR analysis, the enzyme was labeled selectively with [(15)N]alanine. The 13 alanine resonances provide a fingerprint of the protein structure and report on the active site loop conformations and binding of substrate, product, and cofactor. Spectra were recorded for binary and ternary complexes of wild-type DHFR bound to the substrate dihydrofolate (DHF), the product tetrahydrofolate (THF), the pseudosubstrate folate, reduced and oxidized NADPH cofactor, and the inactive cofactor analogue 5,6-dihydroNADPH. The data show that DHFR exists in solution in two dominant conformational states, with the active site loops adopting conformations that closely approximate the occluded or closed conformations identified in earlier X-ray crystallographic analyses. A minor population of a third conformer of unknown structure was observed for the apoenzyme and for the disordered binary complex with 5,6-dihydroNADPH. The reactive Michaelis complex, with both DHF and NADPH bound to the enzyme, could not be studied directly but was modeled by the ternary folate:NADP(+) and dihydrofolate:NADP(+) complexes. From the NMR data, we are able to characterize the active site loop conformation and the occupancy of the substrate and cofactor binding sites in all intermediates formed in the extended catalytic cycle. In the dominant kinetic pathway under steady-state conditions, only the holoenzyme (the binary NADPH complex) and the Michaelis complex adopt the closed loop conformation, and all product complexes are occluded. The catalytic cycle thus involves obligatory conformational transitions between the closed and occluded states. Parallel studies on the catalytically impaired G121V mutant DHFR show that formation of the closed state, in which the nicotinamide ring of the cofactor is inserted into the active site, is energetically disfavored. The G121V mutation, at a position distant from the active site, interferes with coupled loop movements and appears to impair catalysis by destabilizing the closed Michaelis complex and introducing an extra step into the kinetic pathway.  相似文献   

12.
Type II R67 dihydrofolate reductase (DHFR) is a bacterial plasmid-encoded enzyme that is intrinsically resistant to the widely-administered antibiotic trimethoprim. R67 DHFR is genetically and structurally unrelated to E. coli chromosomal DHFR and has an unusual architecture, in that four identical protomers form a single symmetrical active site tunnel that allows only one substrate binding/catalytic event at any given time. As a result, substitution of an active-site residue has as many as four distinct consequences on catalysis, constituting an atypical model of enzyme evolution. Although we previously demonstrated that no single residue of the native active site is indispensable for function, library selection here revealed a strong bias toward maintenance of two native protomers per mutated tetramer. A variety of such “half-native” tetramers were shown to procure native-like catalytic activity, with similar KM values but kcat values 5- to 33-fold lower, illustrating a high tolerance for active-site substitutions. The selected variants showed a reduced thermal stability (Tm ∼12°C lower), which appears to result from looser association of the protomers, but generally showed a marked increase in resilience to heat denaturation, recovering activity to a significantly greater extent than the variant with no active-site substitutions. Our results suggest that the presence of two native protomers in the R67 DHFR tetramer is sufficient to provide native-like catalytic rate and thus ensure cellular proliferation.  相似文献   

13.
We have used an indirect-capture enzyme-linked immunosorbent assay to quantitate the reactivity of sera from human immunodeficiency virus type 1 (HIV-1)-infected humans with native recombinant gp120 (HIV-1 IIIB or SF-2) or with the gp120 molecule (IIIB or SF-2) denatured by being boiled in the presence of dithiothreitol with or without sodium dodecyl sulfate. Denaturation of IIIB gp120 reduced the titers of sera from randomly selected donors by at least 100-fold, suggesting that the majority of cross-reactive anti-gp120 antibodies present are directed against discontinuous or otherwise conformationally sensitive epitopes. When SF-2 gp120 was used, four of eight serum samples reacted significantly with the denatured protein, albeit with ca. 3- to 50-fold reductions in titer. Only those sera reacting with denatured SF-2 gp120 bound significantly to solid-phase-adsorbed SF-2 V3 loop peptide, and none bound to IIIB V3 loop peptide. Almost all antibody binding to reduced SF-2 gp120 was blocked by preincubation with the SF-2 V3 loop peptide, as was about 50% of the binding to native SF-2 gp120. When sera from a laboratory worker or a chimpanzee infected with IIIB were tested, the pattern of reactivity was reversed, i.e., there was significant binding to reduced IIIB gp120, but not to reduced SF-2 gp120. Binding of these sera to reduced IIIB gp120 was 1 to 10% that to native IIIB gp120 and was substantially decreased by preincubation with IIIB (but not SF-2) V3 loop peptide. To analyze which discontinuous or conformational epitopes were predominant in HIV-1-positive sera, we prebound monoclonal antibodies (MAbs) to IIIB gp120 and then added alkaline phosphatase-labelled HIV-1-positive sera. MAbs (such as 15e) that recognize discontinuous epitopes and compete directly with CD4 reduced HIV-1-positive sera binding by about 50%, whereas neutralizing MAbs to the C4, V2, and V3 domains of gp120 were either not inhibitory or only weakly so. Thus, antibodies to the discontinuous CD4-binding site on gp120 are prevalent in HIV-1-positive sera, antibodies to linear epitopes are less common, most of the antibodies to linear epitopes are directed against the V3 region, and most cross-reactive antibodies are directed against discontinuous epitopes, including regions involved in CD4 binding.  相似文献   

14.
R67 dihydrofolate reductase (R67 DHFR) is a novel protein encoded by an R-plasmid that confers resistance to the antibiotic, trimethoprim. This homotetrameric enzyme possesses 222 symmetry, which imposes numerous constraints on the single active site pore, including a "one-site-fits-both" strategy for binding its ligands, dihydrofolate (DHF) and NADPH. Previous studies uncovered salt effects on binding and catalysis (Hicks, S. N., Smiley, R. D., Hamilton, J. B., and Howell, E. E. (2003) Biochemistry 42, 10569-10578), however the one or more residues that participate in ionic contacts with the negatively charged tail of DHF as well as the phosphate groups in NADPH were not identified. Several studies predict that Lys-32 residues were involved, however mutations at this residue destabilize the R67 DHFR homotetramer. To study the role of Lys-32 in binding and catalysis, asymmetric K32M mutations have been utilized. To create asymmetry, individual mutations were added to a tandem array of four in-frame gene copies. These studies show one K32M mutation is tolerated quite well, whereas addition of two mutations has variable effects. Two double mutants, K32M:1+2 and K32M: 1+4, which place the mutations on opposite sides of the pore, reduce kcat. However a third double mutant, K32M: 1+3, that places two mutations on the same half pore, enhances kcat 4- to 5-fold compared with the parent enzyme, albeit at the expense of weaker binding of ligands. Because the kcat/Km values for this double mutant series are similar, these mutations appear to have uncovered some degree of non-productive binding. This non-productive binding mode likely arises from formation of an ionic interaction that must be broken to allow access to the transition state. The K32M:1+3 mutant data suggest this interaction is an ionic interaction between Lys-32 and the charged tail of dihydrofolate. This unusual catalytic scenario arises from the 222 symmetry imposed on the single active site pore.  相似文献   

15.
M K Das  J Lindstrom 《Biochemistry》1991,30(9):2470-2477
Concurrent synthesis of overlapping octameric peptides corresponding to the sequence of the Torpedo acetylcholine receptor (AChR) alpha subunit has been carried out on polypropylene supports functionalized with primary amino groups according to a method developed by M. Geysen [(1987) J. Immunol. Methods 102, 259-274]. The peptides on the solid supports have been used in an enzyme-linked immunosorbent assay. Interactions of the synthetic peptides with antibodies are then detected without removing them from the solid support. By this procedure, epitopes of both antisera and monoclonal antibodies to the Torpedo acetylcholine receptor, its subunits, and synthetic peptide fragments have been mapped. Both rat and rabbit antisera to the alpha subunit show major epitopes spanning the residues 150-165, 338-345, and 355-366 on the Torpedo AChR alpha subunit. Epitopes of monoclonal antibodies to these major epitopes and to others have been rather precisely mapped by using this technique with peptides of varying lengths. The specificity of several of these mAbs are of interest because they have been used in mapping the transmembrane orientation of the AChR alpha-subunit polypeptide chain.  相似文献   

16.
Antiserum raised in chickens to dihydrofolate reductase purified from L1210 leukemia cells by affinity chromatography inhibited the catalytic activity and the binding of methotrexate by the enzyme. Lysates of human chronic myelogenous leukemia cells, which had neither catalytic activity for dihydrofolate reductase nor binding of methotrexate, blocked the inhibiting effect of the antiserum on the function of the enzyme in L1210 cell lysates. In double immunodiffusion, these human leukemia cell lysates formed a single precipitin line against the antiserum. These findings indicate that nonfunctional dihydrofolate reductase in human leukemia cells share an antigenic determinant(s) with a functional form of the enzyme from L1210 murine leukemia cells.  相似文献   

17.
We have examined the pattern of dihydrofolate reductase (DHFR) enzyme and mRNA levels in cell cycle stage-specific populations obtained by centrifugal elutriation in Chinese hamster ovary cells and in a derivative line in which the dihydrofolate reductase gene is amplified approximately 50-fold. On a per cell basis, we observed a 2-fold increase in DHFR activity as cells progressed from G1 to G2/M with a concomitant 2-fold increase in the rate of protein synthesis and steady state level of mRNA. Analysis of DHFR mRNA levels in cell cycle stage-specific mouse 3T6 and human 143 tk- cells gave a similar pattern. We also demonstrate that simple alterations in growth conditions prior to elutriations can dramatically increase the levels of DHFR mRNA in all cell cycle states, thereby indicating that growth response associated with the DHFR gene functions independent of the cell cycle. We conclude that during periods of exponential growth the increases in dihydrofolate reductase activity, rate of protein synthesis, and steady state levels of mRNA parallel the general increases in cell volume and protein content associated with normal progression through the cell cycle, and therefore DHFR cannot be considered a cell cycle-regulated enzyme.  相似文献   

18.
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   

19.
Human dihydrofolate reductase-like 1 (DHFRL1) has been identified as a second human dihydrofolate reductase (DHFR) enzyme. Although DHFRL1 have high sequence homology with human DHFR, dihydrofolate (DHF) exhibits a lowered binding affinity to DHFRL1 and the corresponding molecular mechanism is still unknown. To address this question, we studied the binding of DHF to DHFRL1 and DHFR by using molecular dynamics simulation. Moreover, to investigate the role the 24th residue of DHFR/DHFRL1 plays in DHF binding, R24W DHFRL1 mutant was also studied. The van der Waals interaction are more crucial for the total DHF binding energies, while the difference between the DHF binding energies of human DHFR and DHFRL1 can be attributed to the electrostatic interaction and the polar desolvation free energy. More specifically, lower DHF affinity to DHFRL1 can be mainly attributed to the reduction of net electrostatic interactions of residues Arg32 and Gln35 of DHFRL1 with DHF as being affected by Arg24. The side chain of Arg24 in DHFRL1 can extend deeply into the binding sites of DHF and NADPH, and disturb the DHF binding by steric effect, which rarely happens in human DHFR and R24W DHFRL1 mutant. Additionally, the conformation of loop I in DHFRL1 was also studied in this work. Interestingly, the loop conformation resemble to normal closed state of Escherichia coli DHFR other than the closed state of human DHFR. We hope this work will be useful to understand the general characteristics of DHFRL1.  相似文献   

20.
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   

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