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.
Transferred nuclear Overhauser effect measurements have been made on complexes of NADP+ and thioNADP+ with Lactobacillus casei dihydrofolate reductase to provide information about the glycosidic bond conformations in these complexes. Both NADP+ and thioNADP+ are shown to have very similar anti conformations about their adenine glycosidic bonds when bound to the enzyme. However, their nicotinamide glycosidic bond conformations are very different: while NADP+ binds in an exclusively anti conformation, thioNADP+ binds with a distribution of syn/anti conformations very similar to that observed in nicotinamide mononucleotides in free solution (approximately 50:50). Thus for thioNADP+, binding to the enzyme does not significantly perturb the potential function for rotation about the nicotinamide glycosidic bond. Earlier NMR studies [Hyde, E. I., Birdsall, B., Roberts, G. C. K., Feeney, J., & Burgen, A. S. V. (1980) Biochemistry 19, 3738] had indicated that large downfield 1H shifts of the nicotinamide ring protons (0.61-1.36 ppm) are detected on binding NADP+ while only very small shifts (less than 0.1 ppm) are observed in complexes with thioNADP+. The chemical shift and conformational findings are best explained if the thionicotinamide ring extends into solution making essentially no contacts with the enzyme.  相似文献   

2.
The complex of Lactobacillus casei dihydrofolate reductase with the substrate folate and the coenzyme NADP+ has been shown to exist in solution as a mixture of three slowly interconverting conformations whose proportions are pH-dependent [Birdsall, B., Gronenborn, A. M., Hyde, E. I., Clore, G. M., Roberts, G. C. K., Feeney, J., & Burgen, A. S. V. (1982) Biochemistry 21, 5831]. The assignment of the resonances of all the aromatic protons of the ligand molecules in all three conformational states of the complex has now been completed by using a variety of NMR methods, particularly two-dimensional exchange experiments. The resonances of the nicotinamide protons of the coenzyme and the pteridine 7-proton of the folate have different chemical shifts in the three conformations, in some cases differing by more than 1 ppm. Comparison of the COSY spectra of the complex at low pH (conformation I) and high pH (conformations IIa and IIb) with that of the enzyme-methotrexate-NADP+ complex shows only slight differences in the conformation of the protein. The pattern of chemical shift changes in the ligand and the protein indicates that the structural differences are localized within the active site of the enzyme. Nuclear Overhauser effects (NOEs) are observed between the nicotinamide 5- and 6-protons and the methyl resonance of Thr 45 at both low and high pH, indicating that there is no major movement of the nicotinamide ring. By contrast, NOEs are observed between the pteridine 7-proton and the methyl protons of Leu 19 and Leu 27 in conformations I and IIa but not in conformation IIb.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
The complex of Lactobacillus casei dihydrofolate reductase with trimethoprim and NADP+ exists in solution as a mixture of approximately equal amounts of two slowly interconverting conformational states [Gronenborn, A., Birdsall, B., Hyde, E. I., Roberts, G. C. K., Feeney, J., & Burgen, A. S. V. (1981) Mol. Pharmacol. 20, 145]. These have now been further characterized by multinuclear NMR experiments, and a partial structural model has been proposed. 1H NMR spectra at 500 MHz show that the environments of six of the seven histidine residues differ between the two conformations. The characteristic 1H and 31P chemical shifts of nuclei of the coenzyme in the two conformations of the complex are identical in analogous complexes formed with a number of trimethoprim analogues, indicating that the nature of the two conformations is the same in each case. The pyrophosphate 31P resonances have been assigned to the two conformations, and integration of the 31P spectrum shows that the ratio of conformation I to conformation II varies from 0.4 to 2.3 in the complexes with the various trimethoprim analogues, the ratio for the trimethoprim complex itself being 1.2. Transferred NOE experiments, together with the 1H and 13C chemical shifts, indicate that in conformation II of the complex the nicotinamide ring of the coenzyme has swung away from the enzyme surface into solution; this is made possible by changes in the conformation of the pyrophosphate moiety. In conformation I, by contrast, the nicotinamide ring remains bound to the enzyme. 13C and 15N experiments show that trimethoprim is protonated on N1 in both conformations of the ternary complex. Analysis of the 1H chemical shifts of trimethoprim in terms of ring current effects shows that in conformation I of the ternary complex trimethoprim retains the same conformation as in its binary complex, but 13C, 15N, and 19F [using 2,4-diamino-5-(3,5-dimethoxy-4-fluoro-benzyl)pyrimidine] experiments show that the environment of both the pyrimidine ring and benzyl ring is affected by the proximity of the coenzyme. Less information is available about the conformation of the inhibitor in conformation II of the complex, but its environment is similar to that in the binary enzyme-inhibitor complex. The implications of the existence of these two conformations of the enzyme for understanding cooperativity in binding between NADP+ and trimethoprim are briefly discussed.  相似文献   

4.
The thermodynamics of coenzyme binding to human cytochrome P450 reductase (CPR) and its isolated FAD-binding domain have been studied by isothermal titration calorimetry. Binding of 2',5'-ADP, NADP(+), and H(4)NADP, an isosteric NADPH analogue, is described in terms of the dissociation binding constant (K(d)), the enthalpy (DeltaH(B)) and entropy (TDeltaS(B)) of binding, and the heat capacity change (DeltaC(p)). This systematic approach allowed the effect of coenzyme redox state on binding to CPR to be determined. The recognition and stability of the coenzyme-CPR complex are largely determined by interaction with the adenosine moiety (K(d2)(')(,5)(')(-ADP) = 76 nM), regardless of the redox state of the nicotinamide moiety. Similar heat capacity change (DeltaC(p)) values for 2',5'-ADP (-210 cal mol(-)(1) K(-)(1)), NADP(+) (-230 cal mol(-)(1) K(-)(1)), and H(4)NADP (-220 cal mol(-)(1) K(-)(1)) indicate no significant contribution from the nicotinamide moiety to the binding interaction surface. The coenzyme binding stoichiometry to CPR is 1:1. This result validates a recently proposed one-site kinetic model [Daff, S. (2004) Biochemistry 43, 3929-3932] as opposed to a two-site model previously suggested by us [Gutierrez, A., Lian, L.-Y., Wolf, C. R., Scrutton, N. S., and Roberts, C. G. K. (2001) Biochemistry 40, 1964-1975]. Calorimetric studies in which binding of 2',5'-ADP to CPR (TDeltaS(B) = -13400 +/- 200 cal mol(-)(1), 35 degrees C) was compared with binding of the same ligand to the isolated FAD-binding domain (TDeltaS(B) = -11200 +/- 300 cal mol(-)(1), 35 degrees C) indicate that the number of accessible conformational substates of the protein increases upon 2',5'-ADP binding in the presence of the FMN-binding domain. This pattern was consistently observed along the temperature range that was studied (5-35 degrees C). This contribution of coenzyme binding energy to domain dynamics in CPR agrees with conclusions from previous temperature-jump studies [Gutierrez, A., Paine, M., Wolf, C. R., Scrutton, N. S., and Roberts, G. C. K. (2002) Biochemistry 41, 4626-4637]. A combination of calorimetry and stopped-flow spectrophotometry kinetics experiments showed that this linkage between coenzyme binding energetics and diffusional domain motion impinges directly on the molecular recognition of cytochrome c by CPR. Single-turnover reduction of cytochrome c by CPR (k(max) = 15 s(-)(1), K(d) = 37 microM) is critically coupled to coenzyme binding through ligand-induced motions that enable the FMN-binding domain to overcome a kinetically unproductive conformation. This is remarkable since the FMN-binding domain is not directly involved in coenzyme binding, the NADP(H) binding site being fully contained in the FAD-binding domain. Sequential rapid mixing measurements indicate that harnessing of coenzyme binding energy to the formation of a kinetically productive CPR-cytochrome c complex is a highly synchronized event. The inferred half-time for the decay of this productive conformation (tau(50)) is 330 +/- 70 ms only. Previously proposed structural and kinetic models are discussed in light of these findings.  相似文献   

5.
Kinetic data for the 6'-aminoglycoside-modifying enzyme, AAC(6')-4, also named kanamycin acetyltransferase, have been collected for six aminoglycoside antibiotics (amikacin, gentamicin C1a, kanamycin A, neomycin B, sisomicin, and tobramycin) and three coenzymes (acetyl-CoA, n-propionyl-CoA, and n-butyryl-CoA). The initial velocity pattern using acetyl-CoA favors a ping-pong kinetic mechanism (Kia = -0.34 +/- 0.34 microM), but the pattern using n-propionyl-CoA supports a sequential one (Kia = 2.7 +/- 0.8 microM). Kinetic analyses using alternative substrates confirm the sequential mechanism and, moreover, clearly identify a random order of addition of antibiotic and coenzyme because V/K values of antibiotics varied 40-fold as the identity of the coenzyme was changed and V/K values of coenzyme varied 13-fold as the identity of the antibiotic was changed. One or both sets of values would have remained unchanged if the mechanism were either ordered sequential or ping-pong. Combining these results with structure-activity data which argue for a rapid rate of release of substrates and products relative to the rate of enzymatic turnover (Radika, K., and Northrop, D. B. (1984) Biochemistry 25, in press) establishes the kinetic mechanism as rapid equilibrium random sequential.  相似文献   

6.
Sharkey MA  Gori A  Capone M  Engel PC 《The FEBS journal》2012,279(17):3003-3009
Active-site mutants of glutamate dehydrogenase from Clostridium?symbiosum have been designed and constructed and the effects on coenzyme preference evaluated by detailed kinetic measurements. The triple mutant F238S/P262S/D263K shows complete reversal in coenzyme selectivity from NAD(H) to NADP(H) with retention of high levels of catalytic activity for the new coenzyme. For oxidized coenzymes, k(cat) /K(m) ratios of the wild-type and triple mutant enzyme indicate a shift in preference of approximately 1.6?×?10(7) -fold, from ~?80?000-fold in favour of NAD(+) to ~?200-fold in favour of NADP(+) . For reduced coenzymes the corresponding figure is 1.7?×?10(4) -fold, from ~?1000-fold in favour of NADH to ~?17-fold in favour of NADPH. A fourth mutation (N290G), previously identified as having a potential bearing on coenzyme specificity, did not engender any further shift in preference when incorporated into the triple mutant, despite having a significant effect when expressed as a single mutant.  相似文献   

7.
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   

8.
We have measured the 13C chemical shifts for trimethoprim molecules selectively enriched with 13C at the 2-, 4-, 5-, 6-, and 7-positions and the p-OCH3 position in their complexes with Lactobacillus casei dihydrofolate reductase in the presence and absence of coenzyme analogues. The C2 carbon shifts indicate that the pyrimidine ring is protonated at N1 in all the complexes of trimethoprim with the enzyme and coenzymes and in each case the pyrimidine ring is binding in a similar way to that of the corresponding part of methotrexate in the enzyme-methotrexate complex. The C6 carbon of trimethoprim shows a large upfield shift in all complexes (3.51 to 4.70 ppm) but no shift in the complex of 2,4-diaminopyrimidine with the enzyme: these shifts probably arise from steric interactions between the C1' and C2' carbons and the H6 proton, which approach van der Waals contact in the folded conformation adopted by trimethoprim when bound to the enzyme. The large shift observed for C6 in all complexes indicates that the basic folded conformation is present in all of them. A comparison of the 13C shifts in the enzyme-trimethoprim-NADPH complex with those in the enzyme-trimethoprim binary complex shows substantial changes even for carbons such as C6 and p-OCH3 (0.46 and -0.36 ppm, respectively), which are remote from the coenzyme: these are caused by ligand-induced conformational changes that may involve displacement of the helix containing residues 42-49.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
The binding of folinic acid (5-formyl-5,6,7,8-tetrahydrofolate) to Lactobacillus casei dihydrofolate reductase has been measured. The natural 6S, alpha S diastereoisomer has a binding constant of 1.3 (+/- 0.6) X 10(8) M-1 at pH 6.0, 25 degrees C; the 6R, alpha S diastereoisomer binds approximately 10(4)-fold more weakly. The natural diastereoisomer of folinic acid binds negatively cooperatively with the coenzymes NADP+ and NADPH, binding 3 times more weakly in the presence of NADP+ and 600 times more weakly in the presence of NADPH than to the enzyme alone. Negative cooperativity has been unequivocally distinguished from competition by measurements of coenzyme binding as a function of folinic acid concentration, of the effects of folinic acid on the 1H and 31P chemical shifts of the bound coenzyme, and of the effects of folinic acid on the coenzyme dissociation rate constant. The latter experiments also give evidence for the coexistence of two slowly interconverting conformational forms of the ternary enzyme-coenzyme-folinic acid complex. Small changes in structure of the oxidized coenzymes have substantial effects on the cooperativity with folinic acid, with the thionicotinamide analogue showing positive rather than negative cooperativity. The changes in environment of the bound coenzyme produced by folinic acid, as revealed by 1H and 31P NMR, demonstrate clearly that the negative cooperativity shown by NADP+ and NADPH, respectively, arises by two structurally distinct mechanisms.  相似文献   

10.
The ionization state of aspartate 26 in Lactobacillus casei dihydrofolate reductase has been investigated by selectively labeling the enzyme with [13Cgamma] aspartic acid and measuring the 13C chemical shifts in the apo, folate-enzyme, and dihydrofolate-enzyme complexes. Our results indicate that no aspartate residue has a pKa greater than approximately 4.8 in any of the three complexes studied. The resonance of aspartate 26 in the dihydrofolate-enzyme complex has been assigned by site-directed mutagenesis; aspartate 26 is found to have a pKa value of less than 4 in this complex. Such a low pKa value makes it most unlikely that the ionization of this residue is responsible for the observed pH profile of hydride ion transfer [apparent pKa = 6.0; Andrews, J., Fierke, C. A., Birdsall, B., Ostler, G., Feeney, J., Roberts, G. C. K., and Benkovic, S. J. (1989) Biochemistry 28, 5743-5750]. Furthermore, the downfield chemical shift of the Asp 26 (13)Cgamma resonance in the dihydrofolate-enzyme complex provides experimental evidence that the pteridine ring of dihydrofolate is polarized when bound to the enzyme. We propose that this polarization of dihydrofolate acts as the driving force for protonation of the electron-rich O4 atom which occurs in the presence of NADPH. After this protonation of the substrate, a network of hydrogen bonds between O4, N5 and a bound water molecule facilitates transfer of the proton to N5 and transfer of a hydride ion from NADPH to the C6 atom to complete the reduction process.  相似文献   

11.
The synergism between coenzyme and anion binding to liver alcohol dehydrogenase has been examined by equilibrium measurements and transient-state kinetic methods to characterize electrostatic interactions of coenzymes with ligands which are bound to the catalytic zinc ion of the enzyme subunit. Inorganic anions typically exhibit an at least 200-fold higher affinity for the general anion-binding site than for catalytic zinc on complex formation with free enzyme. Acetate and SCN- interact more strongly with catalytic zinc in the enzyme X NAD+ complex than with the general anion-binding site in free enzyme. CN- shows no significant affinity for the general anion-binding site, but combines to catalytic zinc in the absence as well as the presence of coenzymes. Coordination of CN- to catalytic zinc weakens the binding of NADH by a factor of 50, and tightens the binding of NAD+ to approximately the same extent through interactions which do not include any contributions from covalent adduct formation between CN- and NAD+. These observations provide unambiguous information about the magnitude of electrostatic field effects of coenzymes on anion (e.g. hydroxyl ion) binding to catalytic zinc. They lead to the important inference that coenzyme binding must be strongly affected by ionization of zinc-bound water irrespective of the actual acidity of the latter group. It is concluded on such grounds that the much debated pH dependence of coenzyme binding to liver alcohol dehydrogenase must derive from ionization of zinc-bound water. The assumption that such is not the case leads to the inference that there is no detectable effect of ionization of zinc-bound water on coenzyme binding over the pH range 6-12, a possibility which is definitely excluded by the present results.  相似文献   

12.
Recent studies demonstrated that wheat germ poly(A)-binding protein (PABP) interacted with translation eukaryotic initiation factor (eIF)-iso4G and eIF4B, and these interactions increased the poly(A) binding activity of PABP (Le, H., Tanguay, R. L., Balasta, M. L., Wei, C. C., Browning, K. S., Metz, A. M., Goss, D. J., and Gallie, D. R. (1997) J. Biol. Chem. 272, 16247-16255) and the cap binding activity of eIF-iso4F (Wei, C. C., Balasta, M. L., Ren, J., and Goss, D. J. (1998) Biochemistry 37, 1910-1916). We report here that the interaction between PABP and eIF-iso4G has a substantial effect on the ATPase activity and RNA helicase activity of (eIF4A + eIF4B + eIF-iso4F) complex. ATPase kinetic assays show, in the presence of poly(U), PABP can increase the parameter (k(cat)/K(m)) by 3.5-fold with a 2-fold decrease of K(m) for the (eIF4A + eIF-iso4F) complex. In the presence of globin messenger RNA, the ATPase activity of the complex (eIF4A + eIF-iso4F) was increased 2-fold by the presence of PABP. RNA helicase assays demonstrated that the presence of PABP enhanced the RNA duplex unwinding activity of the initiation factor complex. These results suggest that, in terms of the scanning model of translation initiation, PABP may enhance the mRNA scanning rate of the complex formed by eIF4A, eIF4B, and eIF4F or eIF-(iso)4F and increase the rate of translation.  相似文献   

13.
Rubach JK  Ramaswamy S  Plapp BV 《Biochemistry》2001,40(42):12686-12694
The participation of Val-292 in catalysis by alcohol dehydrogenase and the involvement of dynamics were investigated. Val-292 interacts with the nicotinamide ring of the bound coenzyme and may facilitate hydride transfer. The substitution of Val-292 with Ser (V292S) increases the dissociation constants for the coenzymes (NAD(+) by 50-fold, NADH by 75-fold) and the turnover numbers by 3-7-fold. The V292S enzyme crystallized in the presence of NAD(+) and 2,3,4,5,6-pentafluorobenzyl alcohol has an open conformation similar to the structure of the wild-type apo-enzyme, rather than the closed conformation observed for ternary complexes with wild-type enzyme. The V292S substitution perturbs the conformational equilibrium of the enzyme and decreases the kinetic complexity, which permits study of the hydride transfer step with steady-state kinetics. Eyring plots show that the DeltaH for the oxidation (V(1)) of the protio and deuterio benzyl alcohols is 13 kcal/mol and that the kinetic isotope effect of 4.1 is essentially temperature-independent. Eyring plots for the catalytic efficiency for reduction of benzaldehyde (V(2)/K(p)) with NADH or NADD are distinctly convex, being temperature-dependent from 5 to 25 degrees C and temperature-independent from 25 to 50 degrees C; the kinetic isotope effect of 3.2 for V(2)/K(p) is essentially independent of the temperature. The temperature dependencies and isotope effects for V(1) and V(2)/K(p) are not adequately explained by semiclassical transition state theory and are better explained by hydride transfer occurring through vibrationally assisted tunneling.  相似文献   

14.
Triclosan, a common antibacterial additive used in consumer products, is an inhibitor of FabI, the enoyl reductase enzyme from type II bacterial fatty acid biosynthesis. In agreement with previous studies [Ward, W. H., Holdgate, G. A., Rowsell, S., McLean, E. G., Pauptit, R. A., Clayton, E., Nichols, W. W., Colls, J. G., Minshull, C. A., Jude, D. A., Mistry, A., Timms, D., Camble, R., Hales, N. J., Britton, C. J., and Taylor, I. W. (1999) Biochemistry 38, 12514-12525], we report here that triclosan is a slow, reversible, tight binding inhibitor of the FabI from Escherichia coli. Triclosan binds preferentially to the E.NAD(+) form of the wild-type enzyme with a K(1) value of 23 pM. In agreement with genetic selection experiments [McMurry, L. M., Oethinger, M., and Levy, S. B. (1998) Nature 394, 531-532], the affinity of triclosan for the FabI mutants G93V, M159T, and F203L is substantially reduced, binding preferentially to the E.NAD(+) forms of G93V, M159T, and F203L with K(1) values of 0.2 microM, 4 nM, and 0.9 nM, respectively. Triclosan binding to the E.NADH form of F203L can also be detected and is defined by a K(2) value of 51 nM. We have also characterized the Y156F and A197M mutants to compare and contrast the binding of triclosan to InhA, the homologous enoyl reductase from Mycobacterium tuberculosis. As observed for InhA, Y156F FabI has a decreased affinity for triclosan and the inhibitor binds to both E.NAD(+) and E.NADH forms of the enzyme with K(1) and K(2) values of 3 and 30 nM, respectively. The replacement of A197 with Met has no impact on triclosan affinity, indicating that differences in the sequence of the conserved active site loop cannot explain the 10000-fold difference in affinities of FabI and InhA for triclosan.  相似文献   

15.
NADH is transferred directly from one dehydrogenase enzyme site to another without intervention of the aqueous solvent whenever the two dehydrogenases are of opposite chiral specificity as regards the C4 H of NADH which is transferred in the catalyzed reduction reaction. When both enzymes catalyze the transfer of hydrogen from the same face of the nicotinamide ring, direct enzyme-enzyme transfer of NADH is not possible [Srivastava, D. K., & Bernhard, S. A. (1984) Biochemistry 23, 4538-4545; Srivastava, D. K., & Bernhard, S. A. (1985) Biochemistry (preceding paper in this issue)]. Utilizing an advanced computer graphics facility, and the known three-dimensional coordinates for three dehydrogenases, we have investigated the feasibility of various aspects of the direct transfer of dinucleotide from the site of one enzyme to the site of the other. The facile passage of the coenzyme through the first enzyme site requires an open protein conformation, characteristic of the apoenzyme rather than the holoenzyme structure. Since two dehydrogenases of the same chirality bind coenzyme in the same conformation, the direct transfer of coenzyme from one site to the other is impossible due to the restriction in molecular rotation of the coenzyme in the path of transfer from one binding site to the other; therefore, coenzyme can only be transferred from one dehydrogenase site to another site via the intermediate dissociation of coenzyme into the aqueous milieu. In contrast, when an A dehydrogenase and a B dehydrogenase are juxtaposed, it is stereochemically feasible to transfer the nicotinamide ring from its specific binding site in one enzyme to the site in the other.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The use of affinity chromatography on wheat germ agglutinin columns to partially purify detergent extracts of muscarinic acetylcholine receptor from porcine atria is described. The procedure results in a 20-fold purification of the protein. The partially purified protein binds [3H]L-QNB (the L isomer of quinuclidinyl benzilate) with an observed association rate constant 2- to 3-fold lower than that found for the detergent extract; however, incubation with column fractions eluted prior to the receptor gives an association rate constant similar to that for detergent extracts. The component responsible for this effect is nondialyzable and protease sensitive, indicating that it may be a protein or high-molecular-weight peptide. Affinity labeling experiments with [3H]propylbenzilylcholine mustard [N. J. M. Birdsall, A. S. V. Burgen, and E. C. Hulme (1979) Brit. J. Pharmacol. 66, 337-342] show radioactivity incorporated mainly in a broad peak of apparent molecular weight 75,000 +/- 5000.  相似文献   

17.
S Ramaswamy  D H Park  B V Plapp 《Biochemistry》1999,38(42):13951-13959
When horse liver alcohol dehydrogenase binds coenzyme, a rotation of about 10 degrees brings the catalytic domain closer to the coenzyme binding domain and closes the active site cleft. The conformational change requires that a flexible loop containing residues 293-298 in the coenzyme binding domain rearranges so that the coenzyme and some amino acid residues from the catalytic domain can be accommodated. The change appears to control the rate of dissociation of the coenzyme and to be necessary for installation of the proton relay system. In this study, directed mutagenesis produced the activated Gly293Ala/Pro295Thr enzyme. X-ray crystallography shows that the conformations of both free and complexed forms of the mutated enzyme and wild-type apoenzyme are very similar. Binding of NAD(+) and 2,2, 2-trifluoroethanol do not cause the conformational change, but the nicotinamide ribose moiety and alcohol are not in a fixed position. Although the Gly293Ala and Pro295Thr substitutions do not disturb the apoenzyme structure, molecular modeling shows that the new side chains cannot be accommodated in the closed native holoenzyme complex without steric alterations. The mutated enzyme may be active in the "open" conformation. The turnover numbers with ethanol and acetaldehyde increase 1.5- and 5.5-fold, respectively, and dissociation constants for coenzymes and other kinetic constants increase 40-2,000-fold compared to those of the native enzyme. Substrate deuterium isotope effects on the steady state V or V/K(m) parameters of 4-6 with ethanol or benzyl alcohol indicate that hydrogen transfer is a major rate-limiting step in catalysis. Steady state oxidation of benzyl alcohol is most rapid above a pK of about 9 for V and V/K(m) and is 2-fold faster in D(2)O than in H(2)O. The results are consistent with hydride transfer from a ground state zinc alkoxide that forms a low-barrier hydrogen bond with the hydroxyl group of Ser48.  相似文献   

18.
Analogues of a synthetic heptapeptide substrate corresponding to the sequence around a phosphorylation site in histone H2B [Glass, D. B. & Krebs, E. G. (1982) J. Biol. Chem. 257, 1196-1200] were used to assess interactions between the peptide substrate and the ATP binding sites of cGMP-dependent protein kinase and the catalytic subunit of cAMP-dependent protein kinase. The affinity of each protein kinase for lin-benzo-ADP was determined in the absence and presence of substrate peptide by fluorescence anisotropy titrations [Bhatnagar, D., Roskoski, R., Jr., Rosendahl, M. S., & Leonard, N. J. (1983) Biochemistry 22, 6310-6317]. The Kd values of cGMP-dependent protein kinase for lin-benzo-ADP in the absence and presence of cGMP were 7.6 and 9.7 microM, respectively. Histone H2B(29-35) (Arg-Lys-Arg-Ser-Arg-Lys-Glu) had no effect on nucleotide affinity in either the absence or presence of cGMP. However, when lysine-34 located two residues after the phosphorylatable serine is replaced with an alanyl residue, the resulting [Ala34]histone H2B(29-35) and its analogue peptides interact with cGMP-dependent protein kinase and/or the nucleotide in a fashion that decreases nucleotide binding affinity approximately 3-fold. This amino acid replacement had previously been shown to cause an increase in Vmax and a decrease in the pH optimum for the phosphotransferase reaction. Replacement of positively charged residues at positions 30 and 31 of the peptide also decreased nucleotide affinity. Other analogues of histone H2B(29-35) failed to affect binding of lin-benzo-ADP to the active site of the cGMP-dependent enzyme.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Ferredoxin-NADP(+) reductase catalyses NADP(+) reduction, being specific for NADP(+)/H. To understand coenzyme specificity determinants and coenzyme specificity reversion, mutations at the NADP(+)/H pyrophosphate binding and of the C-terminal regions have been simultaneously introduced in Anabaena FNR. The T155G/A160T/L263P/Y303S mutant was produced. The mutated enzyme presents similar k(cat) values for NADPH and NADH, around 2.5 times slower than that reported for WT FNR with NADPH. Its K(m) value for NADH decreased 20-fold with regard to WT FNR, whereas the K(m) for NADPH remains similar. The combined effect is a much higher catalytic efficiency for NAD(+)/H, with a minor decrease of that for NADP(+)/H. In the mutated enzyme, the specificity for NADPH versus NADH has been decreased from 67,500 times to only 12 times, being unable to discriminate between both coenzymes. Additionally, giving the role stated for the C-terminal Tyr in FNR, its role in the energetics of the FAD binding has been analysed.  相似文献   

20.
The L-21 ScaI ribozyme derived from the intervening sequence of Tetrahymena thermophila pre-rRNA catalyzes a guanosine-dependent endonuclease reaction that is analogous to the first step in self-splicing of this intervening sequence. We now describe pre-steady-state kinetic experiments, with sulfur substituting for the pro-RP (nonbridging) phosphoryl oxygen atom at the site of cleavage, that test aspects of a kinetic model proposed for the ribozyme reaction (Herschlag, D., & Cech, T. R. (1990) Biochemistry 29, 10159-10171). Thio substitution does not affect the reaction with subsaturating oligonucleotide substrate and saturating guanosine ((kcat/Km)S), consistent with the previous finding that binding of the oligonucleotide substrate limits this rate constant. In contrast, there is a significant decrease in the rate of single-turnover reactions of ribozyme-bound (i.e., saturating) oligonucleotide substrate upon thio substitution, with decreases of 2.3-fold for the reaction with guanosine ((kcat/Km)G) and 7-fold for hydrolysis [i.e., with solvent replacing guanosine; kc(-G)]. These "thio effects" are consistent with rate-limiting chemistry, as shown by comparison with model reactions. Nonenzymatic nucleophilic substitution reactions of the phosphate diester, methyl 2,4-dinitrophenyl phosphate monoanion, are slowed 4-11-fold by thio substitution for reactions with hydroxide ion, formate ion, fluoride ion, pyridine, and nicotinamide. In addition, we have confirmed that thio substitution has no effect on the nonenzymatic alkaline cleavage of RNA (Burgers, P. M. J., & Eckstein, F. (1979) Biochemistry 18, 592-596). Considering the strong preference of Mg2+ for binding to oxygen rather than sulfur, the modest thio effect on the chemical step of the ribozyme-catalyzed reaction and the absence of a thio effect on the equilibrium constant for binding of the oligonucleotide substrate suggest that the pro-RP oxygen atom is not coordinated to Mg2+ in the E.S complex or in the transition state. General implications of thio effects in enzymatic reactions of phosphate diesters 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