首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
A site-directed mutagenesis, D244E, of S-adenosylhomocysteine hydrolase (AdoHcyase) changes drastically the nature of the protein, especially the NAD(+) binding affinity. The mutant enzyme contained NADH rather than NAD(+) (Gomi, T., Takata, Y., Date, T., Fujioka, M., Aksamit, R. R., Backlund, P. S., and Cantoni, G. L. (1990) J. Biol. Chem. 265, 16102-16107). In contrast to the site-directed mutagenesis study, the crystal structures of human and rat AdoHcyase recently determined have shown that the carboxyl group of Asp-244 points in a direction opposite to the bound NAD molecule and does not participate in any hydrogen bonds with the NAD molecule. To explain the discrepancy between the mutagenesis study and the x-ray studies, we have determined the crystal structure of the recombinant rat-liver D244E mutant enzyme to 2.8-A resolution. The D244E mutation changes the enzyme structure from the open to the closed conformation by means of a approximately 17 degrees rotation of the individual catalytic domains around the molecular hinge sections. The D244E mutation shifts the catalytic reaction from a reversible to an irreversible fashion. The large affinity difference between NAD(+) and NADH is mainly due to the enzyme conformation, but not to the binding-site geometry; an NAD(+) in the open conformation is readily released from the enzyme, whereas an NADH in the closed conformation is trapped and cannot leave the enzyme. A catalytic mechanism of AdoHcyase has been proposed on the basis of the crystal structures of the wild-type and D244E enzymes.  相似文献   

2.
Endoxylanases are a group of enzymes that hydrolyze the beta-1, 4-linked xylose backbone of xylans. They are predominantly found in two discrete sequence families known as glycoside hydrolase families 10 and 11. The Streptomyces lividans xylanase Xyl10A is a family 10 enzyme, the native structure of which has previously been determined by x-ray crystallography at a 2.6 A resolution (Derewenda, U., Swenson, L., Green, R., Wei, Y., Morosoli, R., Shareck, F., Kluepfel, D., and Derewenda, Z. S. (1994) J. Biol. Chem. 269, 20811-20814). Here, we report the native structure of Xyl10A refined at a resolution of 1.2 A, which reveals many features such as the rare occurrence of a discretely disordered disulfide bond between residues Cys-168 and Cys-201. In order to investigate substrate binding and specificity in glycoside hydrolase family 10, the covalent xylobiosyl enzyme and the covalent cellobiosyl enzyme intermediates of Xyl10A were trapped through the use of appropriate 2-fluoroglycosides. The alpha-linked intermediate with the nucleophile, Glu-236, is in a (4)C(1) chair conformation as previously observed in the family 10 enzyme Cex from Cellulomonas fimi (Notenboom, V., Birsan, C., Warren, R. A. J., Withers, S. G., and Rose, D. R. (1998) Biochemistry 37, 4751-4758). The different interactions of Xyl10A with the xylobiosyl and cellobiosyl moieties, notably conformational changes in the -2 and -1 subsites, together with the observed kinetics on a range of aryl glycosides, shed new light on substrate specificity in glycoside hydrolase family 10.  相似文献   

3.
Flavocytochrome b(2) from Saccharomyces cerevisiae is a l-lactate/cytochrome c oxidoreductase belonging to a large family of 2-hydroxyacid-dependent flavoenzymes. The crystal structure of the enzyme, with pyruvate bound at the active site, has been determined [Xia, Z.-X., and Mathews, F. S. (1990) J. Mol. Biol. 212, 837-863]. The authors indicate that the methyl group of pyruvate is in close contact with Ala198 and Leu230. These two residues are not well-conserved throughout the family of (S)-2-hydroxy acid oxidases/dehydrogenases. Thus, to probe substrate specificity in flavocytochrome b(2), these residues have been substituted by glycine and alanine, respectively. Kinetic studies on the L230A mutant enzyme and the A198G/L230A double mutant enzyme indicate a change in substrate selectivity for the enzyme toward larger (S)-2-hydroxy acids. In particular, the L230A enzyme is more efficient at utilizing (S)-2-hydroxyoctanoate by a factor of 40 as compared to the wild-type enzyme [Daff, S., Manson, F. D. C., Reid, G. A., and Chapman, S. K. (1994) Biochem. J. 301, 829-834], and the A198G/L230A double mutant enzyme is 6-fold more efficient with the aromatic substrate l-mandelate than it is with l-lactate [Sinclair, R., Reid, G. A., and Chapman, S. K. (1998) Biochem. J. 333, 117-120]. To complement these solution studies, we have solved the structure of the A198G/L230A enzyme in complex with pyruvate and as the FMN-sulfite adduct (both to 2.7 A resolution). We have also obtained the structure of the L230A mutant enzyme in complex with phenylglyoxylate (the product of mandelate oxidation) to 3.0 A resolution. These structures reveal the increased active-site volume available for binding larger substrates, while also confirming that the integrity of the interactions important for catalysis is maintained. In addition to this, the mode of binding of the bulky phenylglyoxylate at the active site is in accordance with the operation of a hydride transfer mechanism for substrate oxidation/flavin reduction in flavocytochrome b(2), whereas a mechanism involving the formation of a carbanion intermediate would appear to be sterically prohibited.  相似文献   

4.
A 569-base pair fragment encompassing the upstream regulatory region, the RNA initiation sites, and the initial part of the coding region of the Saccharomyces cerevisiae alcohol dehydrogenase II gene has been analyzed for the presence of sites which undergo conformational modification under torsional stress. Fine mapping of P1 and S1 endonuclease-sensitive sites was obtained on single topoisomers produced by in vitro ligation. It was shown that the upstream activator sequence, the TATA sequence, a region directly upstream to the RNA initiation sites, and several positions in the first segment of the transcribed region change conformation as a function of the applied torsional stress in a precisely coordinate fashion. The superhelical density optima for this coordinate modifications have been determined. Analysis of the conformational changes of the promoter sequence in several naturally occurring (Young, E. T., Williamson, V. M., Taguchi, A., Smith, M., Sledziewski, L., Russel, D., Osterman, J., Denis, C., Cox, D., and Beier, D., (1982) in Genetic Engineering of Microorganisms for Chemicals (Hollander, A., De Moss, R. D., Kaplan, S., Konisky, J., Savage, D., and Wolle, R. S., eds) pp. 335-361, Plenum Publishing Corp., New York) up-promoter constitutive mutants was performed. This analysis has shown that the conformation of functionally relevant sites changes as a function of sequence mutations that have taken place elsewhere; this shows that the conformational behavior of the whole promoter region is linked and suggests transmission in cis of topological effects in RNA polymerase II promoters.  相似文献   

5.
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   

6.
The structures of the major sialylated oligosaccharide chains in swine tracheal mucin glycoprotein were established. The oligosaccharide chains were released by treatment with alkaline borohydride and isolated by gel filtration on Bio-Gel P6 columns and chromatography on DEAE-cellulose. The neutral oligosaccharide chains in this glycoprotein have been characterized in previous studies (Rana, S.S., Chandrasekaran, E.V., Kennedy, J., and Mendicino, J. (1984) J. Biol. Chem. 259, 12899-12907; Chandrasekaran, E.V., Rana, S.S., Davila, M., and Mendicino, J. (1984) J. Biol. Chem. 259, 12908-12914). The present study reports the isolation of four monosialylated chains ranging in length from 6 to 14 sugar units, two disialylated chains containing 6 and 12 sugar units, and one trisialylated chain containing 9 sugar units. The structure of the sialylated oligosaccharides was determined by permethylation analysis and sequential hydrolysis with specific exoglycosidases. The following structures (where GalNAcol is N-acetylgalactosaminitol) were assigned to these oligosaccharides.  相似文献   

7.
alpha-Mannosidase IIx (MX) is an enzyme closely related to alpha-mannosidase II (MII), a key enzyme in N-glycan biosynthesis that catalyzes the first step in conversion of hybrid- to complex-type N-glycans in Golgi apparatus. Recently we generated MII/MX double knock-out mice and found that double nulls completely lack the complex-type N-glycans (Akama, T. O., Nakagawa, H., Wong, N. K., Sutton-Smith, M., Dell, A., Morris, H. R., Nakayama, J., Nishimura, S.-I., Pai, A., Moremen, K. W., Marth, J. D., and Fukuda, M. N. (2006) Essential and mutually compensatory roles of alpha-mannosidase II and alpha-mannosidase IIx in N-glycan processing in vivo in mice. Proc. Natl. Acad. Sci. U. S. A. 103, 8983-8988). In the present study, we determined minor but unusual N-glycan structures found in MII/MX double knock-out mice. We identified such N-glycans by a systematic glycomics approach applying a two-dimensional LC mapping database and matrix-dependent selective fragmentation technique in MALDI-TOF/TOF MS, a highly sensitive and reliable technique that provides specific fragmentations enabling the determination of precise oligosaccharide structures including regioisomers (Kurogochi, M., and Nishimura, S.-I. (2004) Structural characterization of N-glycopeptides by matrix-dependent selective fragmentation of MALDI-TOF/TOF tandem mass spectrometry. Anal. Chem. 76, 6097-6101). Quantitative profiling of all N-glycan structures including minor components from MII/MX nulls, MII nulls, MX nulls, and wild-type mice at embryonic day 15.5 yielded a total of 37 species when structural heterogeneity was reduced by the removal of the sialic acids. Among six unusual N-glycan structures, two glycoforms were novel and were found only in MII/MX double nulls. We characterize such structure as pseudocomplex-type N-glycans. The present study demonstrated that use of the versatile matrix-dependent selective fragmentation method in MALDI-TOF/TOF MS greatly accelerates detailed structural analysis of a trace amount of N-glycans.  相似文献   

8.
We have identified three developmentally regulated oligosaccharide-processing enzyme activities in Dictyostelium discoideum. Two different alpha-mannosidase activities present at extremely low levels in vegetative cells are expressed during development. The first of these activities (MI) rises sharply from 6 to 12 h of development whereas the second activity (MII) rises sharply from 12 to 18 h of development. MI acts on Man9GlcNAc, which it can degrade to Man5GlcNAc but is inactive toward p-nitrophenyl-alpha-D-mannoside (pnpMan). MII acts on pnpMan but not Man9GlcNAc. These activities are distinct from each other and from lysosomal alpha-mannosidase activity as demonstrated by pH optima, substrate specificity, sensitivity to inhibitors and divalent cations, developmental profiles, and solubility. The characteristics of these developmentally regulated alpha-mannosidase activities are similar to those of Golgi alpha-mannosidases I and II from higher eucaryotes, and they appear to catalyze the in vivo formation of processed asparagine-linked oligosaccharides by developed cells. In addition, developed cells have very low levels of a soluble alpha-mannosidase activity, which is the predominant activity in vegetative cells. This soluble vegetative alpha-mannosidase activity has properties that are reminiscent of the endoplasmic reticulum alpha-mannosidase from rat liver. The intersecting N-acetylglucosaminyltransferase activity that we have described recently in vegetative cells of D. discoideum (Sharkey, D. J., and Kornfeld, R. (1989) J. Biol. Chem. 264, 10411-10419) has a developmental profile that is distinct from that of either of the alpha-mannosidase activities. It has maximum activity at 6 h of development and decreases sharply to its minimum level by 12 h of development. The changes that occur in the levels of these three processing enzymes with development correlate well with the different arrays of asparagine-linked oligosaccharides found in early and late stages of development (Sharkey, D. J., and Kornfeld, R. (1991) J. Biol. Chem. 266, 18485-18497).  相似文献   

9.
Hyperargininemia is a rare autosomal disorder that results from a deficiency in hepatic type I arginase. This deficiency is the consequence of random point mutations that occur throughout the gene. The G235R patient mutation has been proposed to affect the catalytic activity and structural integrity of the protein [D. E. Ash, L. R. Scolnick, Z. F. Kanyo, J. G. Vockley, S. D. Cederbaum, and D. W. Christianson (1998) Mol. Genet. Metab. 64, 243-249]. The G235R (patient) and G235A (control) arginase mutants of rat liver arginase have been generated to probe the effects of these point mutations on the structure and function of hepatic type I arginase. Both mutant arginases were trimeric by gel filtration, but the control G235A mutant had 56% of wild-type activity and the G235R mutant had less than 0.03% activity compared to the wild-type enzyme. The G235R mutant contained undetectable levels of tightly bound manganese as determined by electron paramagnetic resonance, while the G235A mutant had a Mn(II) stoichiometry of 2 Mn/subunit. Molecular modeling indicates that the introduction of an arginine residue at position 235 results in a major rearrangement of the metal ligands that compromise Mn(II) binding.  相似文献   

10.
The activities of restriction enzymes are important examples of Mg(II)-dependent hydrolysis of DNA. While a number of crystallographic studies of enzyme-DNA complexes have also involved metal ions, there have been no solution studies exploring the relationship between enzyme conformation and metal-ion binding in restriction enzymes. Using PvuII restriction endonuclease as a model system, we have successfully developed biosynthetic fluorination and NMR spectroscopy as a solution probe of restriction-enzyme conformation. The utility of this method is demonstrated with a study of metal-ion binding by PvuII endonuclease. Replacement of 74% (+/- 10%) of the Tyr residues in PvuII endonuclease by 3-fluorotyrosine produces an enzyme with Mg(II)-supported specific activity and sequence specificity that is indistinguishable from that of the native enzyme. Mn(II) supports residual activity of both the native and fluorinated enzymes; Ca(II) does not support activity in either enzyme, a result consistent with previous studies. 1H- and 19F-NMR spectroscopic studies reveal that while Mg(II) does not alter the enzyme conformation, the paramagnetic Mn(II) produces both short-range spectral broadening and longer range changes in chemical shift. Most interestingly, Ca(II) binding perturbs a larger number of different resonances than Mn(II). Coupled with earlier mutagenesis studies that place Ca(II) in the active site [Nastri, H. G., Evans, P.D., Walker, I.H. & Riggs, P.D. (1997) J. Biol. Chem. 272, 25761-25767], these data suggest that the enzyme makes conformational adjustments to accommodate the distinct geometric preferences of Ca(II) and may play a role in the inability of this metal ion to support activity in restriction enzymes.  相似文献   

11.
The x-ray structure of the unliganded aspartate transcarbamylase reveals that Arg-113 of the catalytic chain is involved in an important set of interactions at the interface between the catalytic and regulatory subunits (Honzatko, R.B., Crawford, J.L., Monaco, H.L., Ladner, J.E., Edwards, B.F.P., Evans, D.R., Warren, S.G., Wiley, D.C., Ladner, R.C., and Lipscomb, W. N. (1982) J. Mol. Biol. 160, 219-263). In order to disturb this interaction, site-directed mutagenesis has been used to replace Arg-113 with glycine. This modification results in a substantial weakening of the interface between the catalytic and regulatory subunits leading to a high tendency for dissociation. The unliganded mutant enzyme exhibits a pH dependence and a sensitivity toward mercurials analogous to that obtained for the relaxed conformation of the wild-type enzyme. Moreover, the presence of saturating concentrations of aspartate is accompanied by only a slight shift in the optimal pH for activity. The bisubstrate analog N-(phosphonacetyl)-L-aspartate induces a 2-fold increase in the sulfhydryl reactivity as compared to the 4-fold increase observed for the wild-type enzyme. Despite this change in the interactions at the interface between the catalytic and regulatory subunits, the mutant enzyme still retains homotropic and heterotropic effects and exhibits a normal affinity for aspartate. Together these data show that a substantial weakening of the catalytic-regulatory interface can occur without altering the allosteric properties of the enzyme. These results also indicate that the intersubunit interactions involving Arg-113, between the polar domain of the catalytic chain and the zinc domain of the regulatory chain, do not participate in the homotropic cooperativity of the enzyme.  相似文献   

12.
Three peptides containing 6-pyridoxyllysine have been isolated from the tryptic digest of the allosteric fructose-1,6-bisphosphate-dependent pyruvate kinase from Escherichia coli, which had been almost completely inactivated with pyridoxal 5'-phosphate. The labelled peptides have been sequenced. The comparison of their sequences with the primary structure of the cat muscle pyruvate kinase allowed to state that peptide I fits the region spanning residues 423-438 (53% identity), peptide II corresponds to residues 442-457 (44% identity) and peptide III encompasses residues 342-368 (70% identity). These findings are discussed in connection with our previous results on the involvement of the three peptides in the catalytic and regulatory properties of the enzyme (Valentini, G., Speranza, M.L., Iadarola, P., Ferri, G. & Malcovati, M. (1988) Biol. Chem. Hoppe-Seyler 369, 1219-1226) and in connection with their location in the three-dimensional structure of the cat muscle pyruvate kinase (Muirhead, H., Clayden, D.A., Lorimer, C.G., Fothergill-Gilmore, L.A., Schiltz, E. & Schmitt, W. (1986) EMBO J. 5, 475-481).  相似文献   

13.
The two preceding papers [Powers, V. M., Koo, C. W., Kenyon, G. L., Gerlt, J. A., & Kozarich, J. W. (1991) Biochemistry (first paper of three in this issue); Neidhart, D. J., Howell, P. L., Petsko, G. A., Powers, V. M., Li, R., Kenyon, G. L., & Gerlt, J. A. (1991) Biochemistry (second paper of three in this issue)] suggest that the active site of mandelate racemase (MR) contains two distinct general acid/base catalysts: Lys 166, which abstracts the alpha-proton from (S)-mandelate, and His 297, which abstracts the alpha-proton from (R)-mandelate. In this paper we report on the properties of the mutant of MR in which His 297 has been converted to asparagine by site-directed mutagenesis (H297N). The structure of H297N, solved by molecular replacement at 2.2-A resolution, reveals that no conformational alterations accompany the substitution. As expected, H297N has no detectable MR activity. However, H297N catalyzes the stereospecific elimination of bromide ion from racemic p-(bromomethyl)mandelate to give p-(methyl)-benzoylformate in 45% yield at a rate equal to that measured for wild-type enzyme; the unreacted p-(bromomethyl)mandelate is recovered as (R)-p-(hydroxymethyl)mandelate. At pD 7.5, H297N catalyzes the stereospecific exchange of the alpha-proton of (S)- but not (R)-mandelate with D2O solvent at a rate 3.3-fold less than that observed for incorporation of solvent deuterium into (S)-mandelate catalyzed by wild-type enzyme. The pD dependence of the rate of the exchange reaction catalyzed by H297N reveals a pKa of 6.4 in D2O, which is assigned to Lys 166.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Two key amino acids, Thr252 and Asp251, are known to be important for dioxygen activation by cytochrome P450cam. We have solved crystal structures of a critical intermediate, the ferrous dioxygen complex (Fe(II)-O2), of the wild-type P450cam and its mutants, D251N and T252A. The wild-type dioxygen complex structure is very much the same as reported previously (Schlichting, I., Berendzen, J., Chu, K., Stock, A. M., Maves, S. A., Benson, D. E., Sweet, R. M., Ringe, D., Petsko, G. A., and Sligar, S. G. (2000) Science 287, 1615-1622) with the exception of higher occupancy and a more ordered structure of the iron-linked dioxygen and two "catalytic" water molecules that form part of a proton relay system to the iron-linked dioxygen. Due to of the altered conformation of the I helix groove these two waters are missing in the D251N dioxygen complex which explains its lower catalytic activity and slower proton transfer to the dioxygen ligand. Similarly, the T252A mutation was expected to disrupt the active site solvent structure leading to hydrogen peroxide formation rather than substrate hydroxylation. Unexpectedly, however, the two "catalytic" waters are retained in the T252A mutant. Based on these findings, we propose that the Thr(252) accepts a hydrogen bond from the hydroperoxy (Fe(III)-OOH) intermediate that promotes the second protonation on the distal oxygen atom, leading to O-O bond cleavage and compound I formation.  相似文献   

15.
L Gross  M F Dunn 《Biochemistry》1992,31(5):1295-1301
The phenol-induced conformational transition in the insulin hexamer is known to involve a large change in structure wherein residues 1-8 of the insulin B-chain are transformed from an extended coil (T-state) to a helix (R-state). This change in protein conformation both exposes a cryptic protein pocket on each subunit to which phenol binds and forces the HisB10 zinc sites to undergo a change in coordination geometry from octahedral to tetrahedral [Derewenda, U., Derewenda, Z., Dodson, E. J., Dodson, G. G., Reynolds, C. D., Smith, G. D., Sparks, C., & Swensen, D. (1989) Nature 338, 593-596]. Substitution of Co(II) for Zn(II) at the HisB10 sites introduces a sensitive chromophoric probe of the structural and chemical events that occur during this allosteric transition [Roy, M., Brader, M. L., Lee, R. W.-K., Kaarsholm, N. C., Hansen, J. F., & Dunn, M. F. (1989) J. Biol. Chem. 264, 19081-19085]. In this study, using rapid-scannig stopped-flow (RSSF) UV-visible spectroscopic studies, we demonstrate that a transient chemical intermediate is formed during the phenol-induced conversion of Co(II)-substituted hexamer from the T-state to the R-state. Decomposition of the RSSF spectra gave a spectrum for the intermediate with d-d transitions consistent with the assignment of the intermediate as either a distorted tetrahedral or a 5-coordinate Co(II) species. Possible structures for the intermediate and the implications of these findings to the allosteric mechanism are considered.  相似文献   

16.
The solution conformation of the DNA duplex d(C1G2C3A4C5L6C7A8C9G10C11).d(G12C13G14T15G16T17G18T19G20C21G22 ) containing the 2'-deoxyribonolactone lesion (L6) in the middle of the sequence has been investigated by NMR spectroscopy and restrained molecular dynamics calculations. Interproton distances have been obtained by complete relaxation matrix analysis of the NOESY cross-peak intensities. These distances, along with torsion angles for sugar rings and additional data derived from canonical A- and B-DNA, have been used for structure refinement by restrained molecular dynamics (rMD). Six rMD simulations have been carried out starting from both regular A- and B-DNA forms. The pairwise rms deviations calculated for each refined structure are <1 A, indicating convergence to essentially the same geometry. The accuracy of the rMD structures has been assessed by complete relaxation matrix back-calculation. The average sixth-root residual index (Rx = 0.052 +/- 0.003) indicated that a good fit between experimental and calculated NOESY spectra has been achieved. Detailed analysis revealed a right-handed DNA conformation for the duplex in which both the T17 nucleotide opposite the abasic site and the lactone ring are located inside the helix. No kinking is observed for this molecule, even at the abasic site step. This structure is compared to that of the oligonucleotide with the identical sequence containing the stable tetrahydrofuran abasic site analogue that we reported previously [Coppel, Y., Berthet, N., Coulombeau, C., Coulombeau, Ce., Garcia, J., and Lhomme, J. (1997) Biochemistry 36, 4817-4830].  相似文献   

17.
S-Adenosylmethionine synthetase (MAT) catalyzes formation of S-adenosylmethionine (SAM) from ATP and l-methionine (Met) and hydrolysis of tripolyphosphate to PP(i) and P(i). Escherichia coli MAT (eMAT) has been crystallized with the ATP analogue AMPPNP and Met, and the crystal structure has been determined at 2.5 A resolution. eMAT is a dimer of dimers and has a 222 symmetry. Each active site contains the products SAM and PPNP. A modeling study indicates that the substrates (AMPPNP and Met) can bind at the same sites as the products, and only a small conformation change of the ribose ring is needed for conversion of the substrates to the products. On the basis of the ternary complex structure and a modeling study, a novel catalytic mechanism of SAM formation is proposed. In the mechanism, neutral His14 acts as an acid to cleave the C5'-O5' bond of ATP while simultaneously a change in the ribose ring conformation from C4'-exo to C3'-endo occurs, and the S of Met makes a nucleophilic attack on the C5' to form SAM. All essential amino acid residues for substrate binding found in eMAT are conserved in the rat liver enzyme, indicating that the bacterial and mammalian enzymes have the same catalytic mechanism. However, a catalytic mechanism proposed recently by González et al. based on the structures of three ternary complexes of rat liver MAT [González, B., Pajares, M. A., Hermoso, J. A., Guillerm, D., Guillerm, G., and Sanz-Aparicio. J. (2003) J. Mol. Biol. 331, 407] is substantially different from our mechanism.  相似文献   

18.
The first discernible intermediate when fully reduced cytochrome c oxidase reacts with O2 is a dioxygen adduct (compound A) of the binuclear heme iron-copper center. The subsequent decay of compound A is associated with transfer of an electron from the low-spin heme a to this center. This reaction eventually produces the ferryl state (F) of this center, but whether an intermediate state may be observed between A and F has been the subject of some controversy. Here we show, using both optical and EPR spectroscopy, that such an intermediate (P(R)) indeed exists and that it exhibits spectroscopic properties quite distinct from F. The optical spectrum of P(R) is similar or identical to the spectrum of the P(M) intermediate that is formed after compound A when two-electron-reduced enzyme reacts with O2. An unusual EPR spectrum with features of a CuB(II) ion that interacts magnetically with a nearby paramagnet [cf. Hansson, O., Karlsson, B., Aasa, R., V?nng?rd, T., and Malmstr?m, B.G (1982) EMBO J. 1, 1295-1297; Blair, D. F., Witt, S. N., and Chan, S. I. (1985) J. Am. Chem. Soc. 107, 7389-7399] can be uniquely assigned to the P(R) intermediate, not being found in either the P(M) or F intermediate. The binuclear center in the P(R) state may be assigned as having an Fe(a3)(IV)=O CuB(II) structure, as in both the P(M) and F states. The spectroscopic differences between these three intermediates are evaluated. The P(R) state has a key role as an initiator of proton translocation by the enzyme, and the thermodynamic and electrostatic bases for this are discussed.  相似文献   

19.
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   

20.
Complete sequence-specific 1H NMR assignments for human insulin   总被引:3,自引:0,他引:3  
A D Kline  R M Justice 《Biochemistry》1990,29(12):2906-2913
Solvent conditions where human insulin could be studied by high-resolution NMR were determined. Both low pH and addition of acetonitrile were required to overcome the protein's self-association and to obtain useful spectra. Two hundred eighty-six 1H resonances were located and assigned to specific sites on the protein by using two-dimensional NMR methods. The presence and position of numerous dNN sequential NOE's indicate that the insulin conformation seen in crystallographic studies is largely retained under these solution conditions. Slowly exchanging protons were observed for seven backbone amide protons and were assigned to positions A15 and A16 and to positions B15-B19. These amides all occur within helical regions of the protein [Chawdhury, S.A., Dodson, E.J., Dodson, G.G., Reynolds, C.D., Tolley, S.P., Blundell, T.L., Cleasby, A., Pitts, J.E., Tickle, I.J., & Wood, S.P. (1983) Diabetologia 25, 460-464].  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