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 531 毫秒
1.
R67 dihydrofolate reductase (DHFR) shares no sequence or structural homology with chromosomal DHFRs. This enzyme arose recently in response to the clinical use of the antibacterial drug trimethoprim. R67 DHFR is a homotetramer possessing a single active site pore. A high-resolution crystal structure shows the homotetramer possesses exact 222 symmetry [Narayana, N., et al. (1995) Nat. Struct. Biol. 2, 1018-1025]. This symmetry dictates four symmetry-related binding sites must exist for each substrate as well as each cofactor. Isothermal titration calorimetry studies, however, indicate only two molecules bind: either two dihydrofolate molecules, two NADPH molecules, or one substrate and one cofactor [Bradrick, T. D., et al. (1996) Biochemistry 35, 11414-11424]. The latter is the productive ternary complex. To evaluate the role of S65, Q67, I68, and Y69 residues, located near the center of the active site pore, site-directed mutagenesis was performed. One mutation in the gene creates four mutations per active site pore which typically result in large cumulative effects. Steady state kinetic data indicate the mutants have altered K(m) values for both cofactor and substrate. For example, the Y69F R67 DHFR displays an 8-fold increase in the K(m) for dihydrofolate and a 20-fold increase in the K(m) for NADPH. Residues involved in ligand binding in R67 DHFR display very little, if any, specificity, consistent with their possessing dual roles in binding. These results support a model where R67 DHFR utilizes an unusual "hot spot" binding surface capable of binding both ligands and indicate this enzyme has adopted a novel yet simple approach to catalysis.  相似文献   

2.
Methenyltetrahydrofolate synthetase (5-formyltetrahydrofolate cyclodehydrase (cyclo-ligase) (ADP-forming) EC 6.3.3.2) catalyzes the ATP- and Mg2+-dependent transformation of 5-formyltetrahydrofolate (leucovorin) to 5,10-methenyltetrahydrofolate. The enzyme has been purified 49,000-fold from human liver by a two-column procedure with Blue Sepharose followed by folinate-Sepharose chromatography. It appears as a single band both on SDS-polyacrylamide gel electrophoresis (Mr 27,000) and on isoelectric focusing (pI = 7.0) and is monomeric, with a molecular weight of 27,000 on gel filtration. Initial-velocity studies suggest that the enzyme catalyzes a sequential mechanism and at 30 degrees C and pH 6.0 the turnover number is 1000 min-1. The enzyme has a higher affinity for its pentaglutamate substrate (Km = 0.6 microM) than for the monoglutamate (Km = 2 microM). The antifolate methotrexate has no inhibitory effect at concentrations up to 350 microM, while methotrexate pentaglutamate is a competitive inhibitor with a Ki = 15 microM. Similarly, dihydrofolate monoglutamate is a weak inhibitor with a Ki = 50 microM, while the pentaglutamate is a potent competitive inhibitor with a Ki of 3.8 microM. Thus, dihydrofolate and methotrexate pentaglutamates could regulate enzyme activity and help explain why leucovorin fails to rescue cells from high concentrations of methotrexate.  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.
Hicks SN  Smiley RD  Hamilton JB  Howell EE 《Biochemistry》2003,42(36):10569-10578
R67 dihydrofolate reductase (DHFR), which catalyzes the NADPH dependent reduction of dihydrofolate to tetrahydrofolate, belongs to a type II family of R-plasmid encoded DHFRs that confer resistance to the antibacterial drug trimethoprim. Crystal structure data reveals this enzyme is a homotetramer that possesses a single active site pore. Only two charged residues in each monomer are located near the pore, K32 and K33. Site-directed mutants were constructed to probe the role of these residues in ligand binding and/or catalysis. As a result of the 222 symmetry of this enzyme, mutagenesis of one residue results in modification at four related sites. All mutants at K32 affected the quaternary structure, producing an inactive dimer. The K33M mutant shows only a 2-4-fold effect on K(m) values. Salt effects on ligand binding and catalysis for K33M and wildtype R67 DHFRs were investigated to determine if these lysines are involved in forming ionic interactions with the negatively charged substrates, dihydrofolate (overall charge of -2) and NADPH (overall charge of -3). Binding studies indicate that two ionic interactions occur between NADPH and R67 DHFR. In contrast, the binding of folate, a poor substrate, to R67 DHFR.NADPH appears weak as a titration in enthalpy is lost at low ionic strength. Steady-state kinetic studies for both wild type (wt) and K33M R67 DHFRs also support a strong electrostatic interaction between NADPH and the enzyme. Interestingly, quantitation of the observed salt effects by measuring the slopes of the log of ionic strength versus the log of k(cat)/K(m) plots indicates that only one ionic interaction is involved in forming the transition state. These data support a model where two ionic interactions are formed between NADPH and symmetry related K32 residues in the ground state. To reach the transition state, an ionic interaction between K32 and the pyrophosphate bridge is broken. This unusual scenario likely arises from the constraints imposed by the 222 symmetry of the enzyme.  相似文献   

5.
A pH-Stat titration method was developed for measuring dihydrofolate reductase (DHFR) activity; this method permits detection of very low DHFR activities corresponding to 100 pmol of substrate reduced per minute. This value is about ten times lower than those observed using the classical spectrophotometric method. This sensitivity makes it possible to measure the DHFR in crude tissue extracts. With beef liver DHFR, Michaelis constants for the cofactor NADPH and the natural substrate determined by this method were 1.9 +/- 0.3 X 10(-5) and 8.5 +/- 0.5 X 10(-7) M, respectively. The inhibition constant of methotrexate, a competitive inhibitor of dihydrofolate, was 3.4 +/- 1.3 X 10(-11) M.  相似文献   

6.
The Bacillus subtilis dihydrofolate reductase (DHFR) gene was expressed in Escherichia coli. The gene product was purified to homogeneity by Butyl-Toyopearl, Toyopearl HW55, and DEAE-Toyopearl column chromatographies, and its molecular properties were compared to those of E. coli DHFR. The specific enzyme activity of the B. subtilis DHFR was 240 units/mg under the standard assay conditions, being about four times higher than that of the E. coli DHFR. Km for coenzyme NADPH was 20.7 microM, a value about three times larger than that of E. coli, whereas Km (1.5 microM) for the substrate, dihydrofolate, was similar to that of E. coli DHFR. This seems to reflect the low homology of the amino acid sequence in residues 61-88 of the two DHFRs where one of the NADPH binding sites is located [Bystrof, C. & Kraut, J. (1991) Biochemistry 30, 2227-2239]. Similar to the E. coli DHFR [Iwakura, M. et al. (1992) J. Biochem. 111, 37-45], the extension of amino acid sequences at the C-terminal end of the B. subtilis DHFR could be attained without loss of the enzyme function or decrease of the protein yield. Thus, the DHFR is useful as a carrier protein for expressing small polypeptides, such as leucine enkephalin, bradykinin, and somatostatin.  相似文献   

7.
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   

8.
9.
Thymidylate synthetase (EC 2.1.1.45) from rat regenerating liver has been purified over 5000-fold to apparent homogeneity by a procedure involving two affinity methods. Molecular weight of the native enzyme was found to be about 68,000, as determined by gel filtration. Electrophoresis in polyacrylamide gels containing sodium dodecyl sulfate yielded a single band of molecular weight of 35,000, suggesting that thymidylate synthetase is a dimer of very similar or identical subunits. The Michaelis constants for deoxyuridylate (dUMP) and (+/-)L-5,10-methylenetetrahydrofolate are 6.8 microM and 65 microM, respectively. Reaction kinetics and product inhibition studies reveal the enzymatic mechanism to be ordered sequential. 5-Fluoro-dUMP, halogenated analog of the nucleotide substrate is a competitive inhibitor of the enzyme, with an apparent Ki value of 5 nM. Amethopterin, analog of the cofactor is also a competitive inhibitor with an apparent Ki value of 23 microM.  相似文献   

10.
Krahn JM  Jackson MR  DeRose EF  Howell EE  London RE 《Biochemistry》2007,46(51):14878-14888
Type II dihydrofolate reductase (DHFR) is a plasmid-encoded enzyme that confers resistance to bacterial DHFR-targeted antifolate drugs. It forms a symmetric homotetramer with a central pore which functions as the active site. Its unusual structure, which results in a promiscuous binding surface that accommodates either the dihydrofolate (DHF) substrate or the NADPH cofactor, has constituted a significant limitation to efforts to understand its substrate specificity and reaction mechanism. We describe here the first structure of a ternary R67 DHFR.DHF.NADP+ catalytic complex, resolved to 1.26 A. This structure provides the first clear picture of how this enzyme, which lacks the active site carboxyl residue that is ubiquitous in Type I DHFRs, is able to function. In the catalytic complex, the polar backbone atoms of two symmetry-related I68 residues provide recognition motifs that interact with the carboxamide on the nicotinamide ring, and the N3-O4 amide function on the pteridine ring. This set of interactions orients the aromatic rings of substrate and cofactor in a relative endo geometry in which the reactive centers are held in close proximity. Additionally, a central, hydrogen-bonded network consisting of two pairs of Y69-Q67-Q67'-Y69' residues provides an unusually tight interface, which appears to serve as a "molecular clamp" holding the substrates in place in an orientation conducive to hydride transfer. In addition to providing the first clear insight regarding how this extremely unusual enzyme is able to function, the structure of the ternary complex provides general insights into how a mutationally challenged enzyme, i.e., an enzyme whose evolution is restricted to four-residues-at-a-time active site mutations, overcomes this fundamental limitation.  相似文献   

11.
Type II dihydrofolate reductases (DHFRs) encoded by the R67 and R388 plasmids are sequence and structurally different from known chromosomal DHFRs. These plasmid-derived DHFRs are responsible for confering trimethoprim resistance to the host strain. A derivative of R388 DHFR, RBG200, has been cloned and its physical properties have been characterized. This enzyme has been shown to transfer the pro-R hydrogen of NADPH to its substrate, dihydrofolate, making it a member of the A-stereospecific class of dehydrogenases [Brito, R. M. M., Reddick, R., Bennett, G. N., Rudolph, F. B., & Rosevear, P. R. (1990) Biochemistry 29,9825]. Two distinct binary RBG200.NADP+ complexes were detected. Addition of NADP+ to RBG200 DHFR results in formation of an initial binary complex, conformation I, which slowly interconverts to a second more stable binary complex, conformation II. The binding of NADP+ to RBG200 DHFR in the second binary complex was found to be weak, KD = 1.9 +/- 0.4 mM. Transferred NOEs were used to determine the conformation of NADP+ bound to RBG200 DHFR. The initial slope of the NOE buildup curves, measured from the intensity of the cross-peaks as a function of the mixing time in NOESY spectra, allowed interproton distances on enzyme-bound NADP+ to be estimated. The experimentally measured distances were used to define upper and lower bound distance constraints between proton pairs in distance geometry calculations. All NADP+ structures consistent with the experimental distance bounds were found to have a syn conformation about the nicotinamide-ribose (X = 94 +/- 26 degrees) and an anti conformation about the adenine-ribose (X = -92 +/- 32 degrees) glycosidic bonds.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
A new form of alcohol dehydrogenase, designated mu-alcohol dehydrogenase, was identified in surgical human stomach mucosa by isoelectric focusing and kinetic determinations. This enzyme was anodic to class I (alpha, beta, gamma) and class II (pi) alcohol dehydrogenases on agarose isoelectric focusing gels. The partially purified mu-alcohol dehydrogenase, specifically using NAD+ as cofactor, catalyzed the oxidation of aliphatic and aromatic alcohols with long chain alcohols being better substrates, indicating a barrel-shape hydrophobic binding pocket for substrate. mu-Alcohol dehydrogenase stood out in high Km values for both ethanol (18 mM) and NAD+ (340 microM) as well as in high Ki value (320 microM) for 4-methylpyrazole, a competitive inhibitor for ethanol. mu-Alcohol dehydrogenase may account for up to 50% of total stomach alcohol dehydrogenase activity and appeared to play a significant role in first-pass metabolism of ethanol in human.  相似文献   

13.
Despite much work, many key aspects of the mechanism of the dihydrofolate reductase (DHFR) catalyzed reduction of dihydrofolate remain unresolved. In bacterial forms of DHFR both substrate and water access to the active site are controlled by the conformation of the mobile M20 loop. In vertebrate DHFRs only one conformation of the residues corresponding to the M20 loop has been observed. Access to the active site was proposed to be controlled by residue 31. MD simulations of chicken DHFR complexed with substrates and cofactor revealed a closing of the side chain of Tyr 31 over the active site on binding of dihydrofolate. This conformational change was dependent on the presence of glutamate on the para-aminobenzoylamide moiety of dihydrofolate. In its absence, the conformation remained open. Although water could enter the active site and hydrogen bond to N5 of dihydrofolate, indicating the feasibility of water as the proton donor, this was not controlled by the conformation of Tyr 31. The water accessibility of the active site was low for both conformations of Tyr 31. However, when hydride was transferred from NADPH to C6 of dihydrofolate before protonation, the average time during which water was found in hydrogen bonding distance to N5 of dihydrofolate in the active site increased almost fivefold. These results indicated that water can serve as the Broensted acid for the protonation of N5 of dihydrofolate during the DHFR catalyzed reduction.  相似文献   

14.
Dihydrofolate reductase (DHFR) from extracts of Mycobacterium smegmatis strain mc2(6) and trimethoprim-resistant mutant mc2(26) was purified to homogeneity. In crude extracts, the specific activity of the enzyme from the trimethoprim resistant strain was comparable to that from the sensitive strain. The DHFR from both sources was purified using affinity chromatography on MTX-Sepharose followed by Mono Q FPLC. The enzyme has an apparent molecular mass of 23 kDa from gel filtration on Sephadex G-100 and from SDS-PAGE. Amino terminal sequence analysis showed homology with DHFRs from a subset of other gram-positive organisms. The purified enzyme from the trimethoprim-sensitive organism exhibited Km values for H2folate and NADPH of 0.68 +/- 0.2 microM and 21 +/- 4 microM, respectively. The Km values for H2folate and NADPH for the enzyme from the drug-resistant organism were 1.8 +/- 0.4 microM and 5.3 +/- 1.5 microM, respectively. A kcat of 4.5 sec-1 was determined for the DHFR from both sources. The enzyme from both sources was competitively inhibited by pyrimethamine and trimethoprim. The Ki value of trimethoprim, for the enzyme from the drug-resistant organism was about six-fold higher than for the enzyme from drug-sensitive strain. Our data suggest that mutation of DHFR contributes to trimethoprim resistance in the mc2(26) strain of M. smegmatis.  相似文献   

15.
A plasmid-encoded dihydrofolate reductase that originated in a clinical isolate of Salmonella typhimurium (phage type 179) moderately resistant to trimethoprim has been isolated and characterized. The dihydrofolate reductase (called type III) was purified to homogeneity using a combination of gel filtration, hydrophobic chromatography, and methotrexate affinity chromatography. Polyacrylamide gel electrophoresis under denaturing and nondenaturing conditions indicated that the enzyme is a 16,900 molecular weight monomeric protein. Kinetic analyses showed that trimethoprim is a relatively tight binding inhibitor (Ki = 19 nM) competitive with dihydrofolate. The enzyme is also extremely sensitive to methotrexate inhibition (Ki = 9 pM) and has a high affinity for dihydrofolate (Km = 0.4 microM). The sequence of the first 20 NH2-terminal residues of the protein shows 50% homology with the trimethoprim-sensitive chromosomal Escherichia coli dihydrofolate reductase and suggests that the two enzymes may be closely related. This is the first example of a plasmid encoding for a monomeric dihydrofolate reductase only moderately resistant to trimethoprim, and a resistance mechanism, dependent in part on the high dihydrofolate affinity of the type III enzyme, is proposed.  相似文献   

16.
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   

17.
M R Hyman  S A Ensign  D J Arp  P W Ludden 《Biochemistry》1989,28(17):6821-6826
Carbonyl sulfide (COS) has been investigated as a rapid-equilibrium inhibitor of CO oxidation by the CO dehydrogenase purified from Rhodospirillum rubrum. The kinetic evidence suggests that the inhibition by COS is largely competitive versus CO (Ki = 2.3 microM) and uncompetitive versus methylviologen as electron acceptor (Ki = 15.8 microM). The data are compatible with a ping-pong mechanism for CO oxidation and COS inhibition. Unlike the substrate CO, COS does not reduce the iron-sulfur centers of dye-oxidized CO dehydrogenase and thus is not an alternative substrate for the enzyme. However, like CO, COS is capable of protecting CO dehydrogenase from slow-binding inhibition by cyanide. A true binding constant (KD) of 2.2 microM for COS has been derived on the basis of the saturable nature of COS protection against cyanide inhibition. The ability of CO, CO2, COS, and related CO/CO2 analogues to reverse cyanide inhibition of CO dehydrogenase is also demonstrated. The kinetic results are interpreted in terms of two binding sites for CO on CO dehydrogenase from R. rubrum.  相似文献   

18.
Beta-carboline-benzoquinolizidine plant alkaloid deoxytubulosine (DTB) was evaluated and assessed for the first time for its biochemical and biological activity employing the biomarker dihydrofolate reductase (DHFR) (5,6,7,8-tetrahydrofolate: NADP+ oxidoreductase, EC 1.5.1.3) as the probe enzyme, a key target in cancer chemotherapy. DHFR, employed in the present investigations was purified from Lactobacillus leichmannii. DTB, isolated from the Indian medicinal plant Alangium lamarckii was demonstrated to exhibit potent cytotoxicity. The alkaloid potently inhibited the cell growth of L. leichmannii and the cellular enzyme activity of DHFR (IC50=40 and 30 microM for the cell growth and enzyme inhibitions, respectively). DTB concentrations >75 microM resulted in a total loss of the DHFR activity, thus suggesting that the beta-carboline-benzoquinolizidine plant alkaloid is a promising potential antitumor agent. Our results are also suggestive of its potential antimicrobial activity. DTB binding to DHFR appears to be slow and reversible. Inhibition kinetics revealed that DHFR has a Ki value of 5x10(-6) M for DTB and that the enzyme inhibition is a simple linear 'non-competitive' type.  相似文献   

19.
Onion and garlic essential oils were previously shown to inhibit mouse skin tumor promotion, as were the enzymes, lipoxygenase, and cyclooxygenase. In the present study, the inhibition of soybean lipoxygenase (EC 1.13.11.12) by onion and garlic components and related compounds was investigated. The IC50 values as well as the kinetic inhibition constants were determined for the most active compounds. Di-(1-propenyl) sulfide, an analog of the substrate moiety required for oxygenase action, was the only irreversible inhibitor observed with Ki = 59 microM and k3 = 0.53/min. Inhibition in the presence of substrate was uncompetitive at 88 and 132 microM linoleic acid with Ki = 129 microM. At 173 microM linoleic acid, however, inhibition was competitive with Ki = 66 microM. Dially trisulfide, allyl methyl trisulfide, and diallyl disulfide were competitive inhibitors, while 1-propenylpropyl sulfide and (E, Z)-4,5,9-trithiadodeca-1,6,11-triene 9-oxide (ajoene) were mixed inhibitors. Nordihydroguaiaretic acid (NDGA), the most potent lipoxygenase inhibitor, was a competitive inhibitor with Ki = 0.29 microM. The results indicate a relative potency of inhibition for structural features in the following order: di(1-propenyl) sulfide greater than an alkenyl trisulfide greater than an alkenyl disulfide. Di(n-propyl) disulfide, a major onion oil component, inhibited neither lipoxygenase nor promotion. Di(1-propenyl) sulfide and ajoene inhibited both. This suggests that the inhibition of lipoxygenase may be involved in antipromotion.  相似文献   

20.
Glucose-6-phosphate dehydrogenase from sporangiophores of Phycomyces blakesleeanus NRRL 1555 (-) was partially purified. The enzyme showed a molecular weight of 85 700 as determined by gel-filtration. NADP+ protected the enzyme from inactivation. Magnesium ions did not affect the enzyme activity. Glucose-6-phosphate dehydrogenase was specific for NADP+ as coenzyme. The reaction rates were hyperbolic functions of substrate and coenzyme concentrations. The Km values for NADP+ and glucose 6-phosphate were 39.8 and 154.4 microM, respectively. The kinetic patterns, with respect to coenzyme and substrate, indicated a sequential mechanism. NADPH was a competitive inhibitor with respect to NADP+ (Ki = 45.5 microM) and a non-competitive inhibitor with respect to glucose 6-phosphate. ATP inhibited the activity of glucose-6-phosphate dehydrogenase. The inhibition was of the linear-mixed type with respect to NADP+, the dissociation constant of the enzyme-ATP complex being 2.6 mM, and the enzyme-NADP+-ATP dissociation constant 12.8 mM.  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