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.
R R Poyner  G H Reed 《Biochemistry》1992,31(31):7166-7173
Phosphonoacetohydroxamate (PhAH) is a tight-binding (Ki = 15 pM) inhibitor of enolase that is believed to mimic the aci-carboxylate form of the intermediate carbanion in the reaction [Anderson, V. E., Weiss, P. M., & Cleland, W. W. (1984) Biochemistry 23, 2779]. Electron paramagnetic resonance (EPR) spectroscopy of Mn2+ has been used to map sites of interaction of PhAH with the two divalent cations at the active site of enolase from bakers' yeast. EPR spectra of enolase-PhAH complexes containing two Mn2+ bound at the active site contain multiple fine structure transitions each with a 45-G 55Mn hyperfine spacing that is a characteristic of spin exchange coupled pairs of Mn2+. Magnetically dilute complexes were obtained by preparation of specific Mg2+/Mn2+ hybrid complexes by manipulating the order of addition of the divalent metal species. Thus, Mn2+ was placed in the higher affinity site by addition of 1 equiv of Mn2+ to a solution of enolase and PhAH, followed by addition of 1 equiv of Mg2+. Reversing the order of addition of Mg2+ and Mn2+ placed Mn2+ in the lower affinity site. Regiospecifically 17O-labeled forms of PhAH were prepared, and the binding of the functional groups on PhAH to Mn2+ at the two metal ion sites was determined from the presence or absence of 17O superhyperfine coupling in the EPR signals. The hydroxamate oxygen is a ligand of Mn2+ at the higher affinity site, a phosphonate oxygen is a ligand of Mn2+ at the lower affinity site, and the carbonyl oxygen is a mu-O bridge of the two metal ions.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Sims PA  Larsen TM  Poyner RR  Cleland WW  Reed GH 《Biochemistry》2003,42(27):8298-8306
The pH dependence of enolase catalysis was studied to understand how enolase is able to utilize both general acid and general base catalysis in each direction of the reaction at near-neutral pHs. Wild-type enolase from yeast was assayed in the dehydration reaction (2-phospho-D-glycerate --> phosphoenolpyruvate + H(2)O) at different pHs. E211Q, a site-specific variant of enolase that catalyzes the exchange of the alpha-proton of 2-phospho-D-glycerate but not the complete dehydration, was assayed in a (1)H/(2)H exchange reaction at different pDs. Additionally, crystal structures of E211Q and E168Q were obtained at 2.0 and 1.8 A resolution, respectively. Analysis of the pH profile of k(cat)/K(Mg) for wild-type enolase yielded macroscopic pK(a) estimates of 7.4 +/- 0.3 and 9.0 +/- 0.3, while the results of the pD profile of the exchange reaction of E211Q led to a pK(a) estimate of 9.5 +/- 0.1. These values permit estimates of the four microscopic pK(a)s that describe the four relevant protonation states of the acid/base catalytic groups in the active site. The analysis indicates that the dehydration reaction is catalyzed by a small fraction of enzyme that is reverse-protonated (i.e., Lys345-NH(2), Glu211-COOH), whereas the hydration reaction is catalyzed by a larger fraction of the enzyme that is typically protonated (i.e., Lys345-NH(3)(+), Glu211-COO(-)). These two forms of the enzyme coexist in a constant, pH-independent ratio. The structures of E211Q and E168Q both show virtually identical folds and active-site architectures (as compared to wild-type enolase) and thus provide additional support to the conclusions reported herein. Other enzymes that require both general acid and general base catalysis likely require reverse protonation of catalytic groups in one direction of the reaction.  相似文献   

3.
Escherichia coli RNase HI has two Mn(2+)-binding sites. Site 1 is formed by Asp10, Glu48, and Asp70, and site 2 is formed by Asp10 and Asp134. Site 1 and site 2 have been proposed to be an activation site and an attenuation site, respectively. However, Glu48 and Asp134 are dispensable for Mn(2+)-dependent activity. In order to identify the Mn(2+)-binding sites of the mutant proteins at Glu48 and/or Asp134, the crystal structures of the mutant proteins E48A-RNase HI*, D134A-RNase HI*, and E48A/D134N-RNase HI* in complex with Mn(2+) were determined. In E48A-RNase HI*, Glu48 and Lys87 are replaced by Ala. In D134A-RNase HI*, Asp134 and Lys87 are replaced by Ala. In E48A/D134N-RNase HI*, Glu48 and Lys87 are replaced by Ala and Asp134 is replaced by Asn. All crystals had two or four protein molecules per asymmetric unit and at least two of which had detectable manganese ions. These structures indicated that only one manganese ion binds to the various positions around the center of the active-site pocket. These positions are different from one another, but none of them is similar to site 1. The temperature factors of these manganese ions were considerably larger than those of the surrounding residues. These results suggest that the first manganese ion required for activation of the wild-type protein fluctuates among various positions around the center of the active-site pockets. We propose that this fluctuation is responsible for efficient hydrolysis of the substrates by the protein (metal fluctuation model). The binding position of the first manganese ion is probably forced to shift to site 1 or site 2 upon binding of the second manganese ion.  相似文献   

4.
Counterions are required for RNA folding, and divalent metal ions such as Mg(2+) are often critical. To dissect the role of counterions, we have compared global and local folding of wild-type and mutant variants of P4-P6 RNA derived from the Tetrahymena group I ribozyme in monovalent and in divalent metal ions. A remarkably simple picture of the folding thermodynamics emerges. The equilibrium folding pathway in monovalent ions displays two phases. In the first phase, RNA molecules that are initially in an extended conformation enforced by charge-charge repulsion are relaxed by electrostatic screening to a state with increased flexibility but without formation of long-range tertiary contacts. At higher concentrations of monovalent ions, a state that is nearly identical to the native folded state in the presence of Mg(2+) is formed, with tertiary contacts that involve base and backbone interactions but without the subset of interactions that involve specific divalent metal ion-binding sites. The folding model derived from these and previous results provides a robust framework for understanding the equilibrium and kinetic folding of RNA.  相似文献   

5.
R R Poyner  W W Cleland  G H Reed 《Biochemistry》2001,40(27):8009-8017
Spectroscopic and kinetic methods have been used to explore the roles of divalent metal ions in the enolase-catalyzed dehydration of 2-phosphoglycerate (2-PGA). Enolase requires 2 equiv of metal ion per active site for maximal activity. Previous crystallographic studies [Larsen, T. M., Wedekind, J. E., Rayment, I., and Reed, G. H. (1996) Biochemistry 35, 4349-4358] showed that both magnesium ions coordinated to the carboxylate group of the substrate/product-a scheme consistent with metal ion assistance in formation of the enolate intermediate. Electron paramagnetic resonance (EPR) data with 17O-labeled forms of phosphoenolpyruvate show that Mn(2+), bound at the lower affinity site, coordinates to one carboxylate oxygen and one phosphate oxygen of the substrate. These observations are fully consistent with the crystallographic data. Plots of activity versus log [metal ion] are bell-shaped, and the inhibitory phases of the profiles have been previously attributed to binding of metal ions at ancillary sites on the enzyme. However, the activation profiles and measurements of 2H kinetic isotope effects support an ordered kinetic mechanism wherein binding of 2-PGA precedes binding of the second metal ion, and release of the second metal ion occurs prior to departure of phosphoenolpyruvate. High concentrations of metal ion lead to inhibition in the ordered mechanism by interfering with product release. The 2H kinetic isotope effect is diminished in the inhibitory phases of the metal ion activation profiles in a manner that is consistent with the predominantly ordered mechanism. Zn(2+) gives lower maximal activity than Mg(2+), apparently due to slow release of Zn(2+) from the product complex. Addition of imidazole increases the maximal rate apparently by accelerating the release of Zn(2+) from the enzyme.  相似文献   

6.
Tinsley RA  Harris DA  Walter NG 《Biochemistry》2004,43(28):8935-8945
The ability of divalent metal ions to participate in both structure formation and catalytic chemistry of RNA enzymes (ribozymes) has made it difficult to separate their cause and effect in ribozyme function. For example, the recently solved crystal structures of precursor and product forms of the cis-cleaving genomic hepatitis delta virus (HDV) ribozyme show a divalent metal ion bound in the active site that is released upon catalysis due to an RNA conformational change. This conformational switch is associated with a repositioning of the catalytically involved base C75 in the active-site cleft, thus controlling catalysis. These findings confirm previous data from fluorescence resonance energy transfer (FRET) on a trans-acting form of the HDV ribozyme that found a global conformational change to accompany catalysis. Here, we further test the conformational switch model by measuring the Mg(2+) dependence of the global conformational change of the trans-acting HDV ribozyme, using circular dichroism and time-resolved FRET as complementary probes of secondary and tertiary structure formation, respectively. We observe significant differences in both structure and Mg(2+) affinity of the precursor and product forms, in the presence and absence of 300 mM Na(+) background. The precursor shortens while the product extends with increasing Mg(2+) concentration, essentially amplifying the structural differences observed in the crystal structures. In addition, the precursor has an approximately 2-fold and approximately 13-fold lower Mg(2+) affinity than the product in secondary and tertiary structure formation, respectively. We also have compared the C75 wild-type with the catalytically inactive C75U mutant and find significant differences in global structure and Mg(2+) affinity for both their precursor and product forms. Significantly, the Mg(2+) affinity of the C75 wild-type is 1.7-2.1-fold lower than that of the C75U mutant, in accord with the notion that C75 is essential for a catalytic conformational change that leads to a decrease in the local divalent metal ion affinity and release of a catalytic metal. Thus, a consistent picture emerges in which divalent metal ions and RNA functional groups are intimately intertwined in affecting structural dynamics and catalysis in the HDV ribozyme.  相似文献   

7.
The structure- and strand-specific phosphodiesterase flap endonuclease-1 (FEN1), the prototypical 5′-nuclease, catalyzes the essential removal of 5′-single-stranded flaps during replication and repair. FEN1 achieves this by selectively catalyzing hydrolysis one nucleotide into the duplex region of substrates, always targeting the 5′-strand. This specificity is proposed to arise by unpairing the 5′-end of duplex to permit the scissile phosphate diester to contact catalytic divalent metal ions. Providing the first direct evidence for this, we detected changes induced by human FEN1 (hFEN1) in the low-energy CD spectra and fluorescence lifetimes of 2-aminopurine in substrates and products that were indicative of unpairing. Divalent metal ions were essential for unpairing. However, although 5′-nuclease superfamily-conserved active-site residues K93 and R100 were required to produce unpaired product, they were not necessary to unpair substrates. Nevertheless, a unique arrangement of protein residues around the unpaired DNA was detected only with wild-type protein, suggesting a cooperative assembly of active-site residues that may be triggered by unpaired DNA. The general principles of FEN1 strand and reaction-site selection, which depend on the ability of juxtaposed divalent metal ions to unpair the end of duplex DNA, may also apply more widely to other structure- and strand-specific nucleases.  相似文献   

8.
Methylaspartate ammonia-lyase (3-methylaspartase, MAL; EC ) catalyzes the reversible anti elimination of ammonia from L-threo-(2S,3S)-3-methylaspartic acid to give mesaconic acid. This reaction lies on the main catabolic pathway for glutamate in Clostridium tetanomorphum. MAL requires monovalent and divalent cation cofactors for full catalytic activity. The enzyme has attracted interest because of its potential use as a biocatalyst. The structure of C. tetanomorphum MAL has been solved to 1.9-A resolution by the single-wavelength anomalous diffraction method. A divalent metal ion complex of the protein has also been determined. MAL is a homodimer with each monomer consisting of two domains. One is an alpha/beta-barrel, and the other smaller domain is mainly beta-strands. The smaller domain partially occludes the C terminus of the barrel and forms a large cleft. The structure identifies MAL as belonging to the enolase superfamily of enzymes. The metal ion site is located in a large cleft between the domains. Potential active site residues have been identified based on a combination of their proximity to a metal ion site, molecular modeling, and sequence homology. In common with all members of the enolase superfamily, the carboxylic acid of the substrate is co-ordinated by the metal ions, and a proton adjacent to a carboxylic acid group of the substrate is abstracted by a base. In MAL, it appears that Lys(331) removes the alpha-proton of methylaspartic acid. This motif is the defining mechanistic characteristic of the enolase superfamily of which all have a common fold. The degree of structural conservation is remarkable given only four residues are absolutely conserved.  相似文献   

9.
The d-mannonate dehydratase (ManD) function was assigned to a group of orthologous proteins in the mechanistically diverse enolase superfamily by screening a library of acid sugars. Structures of the wild type ManD from Novosphingobium aromaticivorans were determined at pH 7.5 in the presence of Mg2+ and also in the presence of Mg2+ and the 2-keto-3-keto-d-gluconate dehydration product; the structure of the catalytically active K271E mutant was determined at pH 5.5 in the presence of the d-mannonate substrate. As previously observed in the structures of other members of the enolase superfamily, ManD contains two domains, an N-terminal alpha+beta capping domain and a (beta/alpha)7beta-barrel domain. The barrel domain contains the ligands for the essential Mg2+, Asp 210, Glu 236, and Glu 262, at the ends of the third, fourth, and fifth beta-strands of the barrel domain, respectively. However, the barrel domain lacks both the Lys acid/base catalyst at the end of the second beta-strand and the His-Asp dyad acid/base catalyst at the ends of the seventh and sixth beta-strands, respectively, that are found in many members of the superfamily. Instead, a hydrogen-bonded dyad of Tyr 159 in a loop following the second beta-strand and Arg 147 at the end of the second beta-strand are positioned to initiate the reaction by abstraction of the 2-proton. Both Tyr 159 and His 212, at the end of the third beta-strand, are positioned to facilitate both syn-dehydration and ketonization of the resulting enol intermediate to yield the 2-keto-3-keto-d-gluconate product with the observed retention of configuration. The identities and locations of these acid/base catalysts as well as of cationic amino acid residues that stabilize the enolate anion intermediate define a new structural strategy for catalysis (subgroup) in the mechanistically diverse enolase superfamily. With these differences, we provide additional evidence that the ligands for the essential Mg2+ are the only conserved residues in the enolase superfamily, establishing the primary functional importance of the Mg2+-assisted strategy for stabilizing the enolate anion intermediate.  相似文献   

10.
L Lebioda  B Stec 《Biochemistry》1991,30(11):2817-2822
Enolase in the presence of Mg2+ catalyzes the elimination of H2O from 2-phosphoglyceric acid (PGA) to form phosphoenolpyruvate (PEP) and the reverse reaction, the hydration of PEP to PGA. The structure of the ternary complex yeast enolase-Mg2(+)-PGA/PEP has been determined by X-ray diffraction and refined by crystallographic restrained least-squares to an R = 16.9% for those data with I/sigma (I) greater than or equal to 2 to 2.2-A resolution with a good geometry of the model. The structure indicates the substrate molecule in the active site has its hydroxyl group coordinated to the Mg2+ ion. The carboxylic group interacts with the side chains of His373 and Lys396. The phosphate group is H-bonded to the guanidinium group of Arg374. A water molecule H-bonded to the carboxylic groups of Glu168 and Glu211 is located at a 2.6-A distance from carbon-2 of the substrate in the direction of its proton. We propose that this cluster functions as the base abstracting the proton in the catalytic process. The proton is probably transferred, first to the water molecule, then to Glu168, and further to the substrate hydroxyl to form a water molecule. Some analogy is apparent between the initial stages of the enolase reverse reaction, the hydration of PEP, and the proteolytic mechanism of the metallohydrolases carboxypeptidase A and thermolysin. The substrate/product binding is accompanied by large movements of loops Ser36-His43 and Ser158-Gly162. The role of these conformational changes is not clear at this time.  相似文献   

11.
12.
Wang TY  Chen YC  Kao LW  Chang CY  Wang YK  Liu YH  Feng JM  Wu TK 《The FEBS journal》2008,275(20):5007-5020
The biofilm-related and carnosine-hydrolyzing aminoacylhistidine dipeptidase (pepD) gene from Vibrio alginolyticus was cloned and sequenced. The recombinant PepD protein was produced and biochemically characterized and the putative active-site residues responsible for metal binding and catalysis were identified. The recombinant enzyme, which was identified as a homodimeric dipeptidase in solution, exhibited broad substrate specificity for Xaa-His and His-Xaa dipeptides, with the highest activity for the His-His dipeptide. Sequence and structural homologies suggest that the enzyme is a member of the metal-dependent metallopeptidase family. Indeed, the purified enzyme contains two zinc ions per monomer. Reconstitution of His.Tag-cleaved native apo-PepD with various metal ions indicated that enzymatic activity could be optimally restored when Zn2+ was replaced with other divalent metal ions, including Mn2+, Co2+, Ni2+, Cu2+ and Cd2+, and partially restored when Zn2+ was replaced with Mg2+. Structural homology modeling of PepD also revealed a 'catalytic domain' and a 'lid domain' similar to those of the Lactobacillus delbrueckii PepV protein. Mutational analysis of the putative active-site residues supported the involvement of His80, Asp119, Glu150, Asp173 and His461 in metal binding and Asp82 and Glu149 in catalysis. In addition, individual substitution of Glu149 and Glu150 with aspartic acid resulted in the partial retention of enzymatic activity, indicating a functional role for these residues on the catalysis and zinc ions, respectively. These effects may be necessary either for the activation of the catalytic water molecule or for the stabilization of the substrate-enzyme tetrahedral intermediate. Taken together, these results may facilitate the design of PepD inhibitors for application in antimicrobial treatment and antibody-directed enzyme prodrug therapy.  相似文献   

13.
In this article, we report the results of an analysis of the glycolytic enzyme enolase (2-phospho-d-glycerate hydrolase) of Trypanosoma brucei. Enolase activity was detected in both bloodstream-form and procyclic insect-stage trypanosomes, although a 4.5-fold lower specific activity was found in the cultured procyclic homogenate. Subcellular localization analysis showed that the enzyme is only present in the cytosol. The T. brucei enolase was expressed in Escherichia coli and purified to homogeneity. The kinetic properties of the bacterially expressed enzyme showed strong similarity to those values found for the natural T. brucei enolase present in a cytosolic cell fraction, indicating a proper folding of the enzyme in E. coli. The kinetic properties of T. brucei enolase were also studied in comparison with enolase from rabbit muscle and Saccharomyces cerevisiae. Functionally, similarities were found to exist between the three enzymes: the Michaelis constant (Km) and KA values for the substrates and Mg2+ are very similar. Differences in pH optima for activity, inhibition by excess Mg2+ and susceptibilities to monovalent ions showed that the T. brucei enolase behaves more like the yeast enzyme. Alignment of the amino acid sequences of T. brucei enolase and other eukaryotic and prokaryotic enolases showed that most residues involved in the binding of its ligands are well conserved. Structure modelling of the T. brucei enzyme using the available S. cerevisiae structures as templates indicated that there are some atypical residues (one Lys and two Cys) close to the T. brucei active site. As these residues are absent from the human host enolase and are therefore potentially interesting for drug design, we initiated attempts to determine the three-dimensional structure. T. brucei enolase crystals diffracting at 2.3 A resolution were obtained and will permit us to pursue the determination of structure.  相似文献   

14.
A A Ismail  H H Mantsch 《Biopolymers》1992,32(9):1181-1186
The carboxylate-containing homopolypeptides poly(L-glutamate) [poly(Glu)] and poly(L-aspartate) [poly(Asp)] were found to form different types of ordered structures in the presence of poly(L-lysine) [poly(Lys)]. Mixing poly(Glu) with poly(Lys) in aqueous solution at neutral pH results in the instantaneous formation of a gel-like precipitate. The secondary structure of the gel precipitate can be best described as intermolecular antiparallel beta-strands, involving the backbone amide groups, as evidenced by the presence of characteristic amide I bands in the ir spectrum at 1684 and 1612 cm-1. Mixing poly(Asp) with poly(Lys) under identical conditions results in the formation of a fine precipitate with a different morphology. Examination of the ir spectrum of the precipitate revealed that unlike poly(Glu), poly(Asp) did not yield any discrete secondary structure upon precipitation with poly(Lys). Addition of solutions containing Ca2+ or Mg2+ to the poly(Glu)/poly(Lys) aggregates resulted in complete dissolution of the gel, with the disappearance of the ir bands characteristic of the intermolecular hydrogen-bonded network. The results demonstrate the importance of salt bridges in establishing strong hydrogen bonds between the backbone amide groups. Reaggregation occurred upon heating the poly(Glu)/poly(Lys) mixture in the presence of Ca2+, but not in the presence of Mg2+ ions. In the presence of Ca2+ ions, aggregation and formation of an extended hydrogen-bonded network occurred upon heating. The aggregates formed upon heating poly(Glu)/poly(Lys) in the presence of Ca2+ were attributed solely to complexation of Ca2+ to the carboxylate groups of poly(Glu) with poly(Lys) remaining free in solution. Dissolution of the aggregate could be accomplished through addition of Mg2+ at room temperature.  相似文献   

15.
The conformational equilibria of the A subunit of DNA gyrase (GyrA), of its 59 kDa N-terminal fragment (GyrA59) and of the quinolone-resistant Ser-Trp83 mutant (GyrATrp83), were investigated in the presence of mono- and divalent metal ions and ciprofloxacin, a clinically useful antibacterial quinolone. The stability of the proteins was estimated from temperature denaturation, monitoring unfolding with circular dichroism spectroscopy. Two transitions were observed in GyrA and GyrATrp83, which likely reflect unfolding of the N and C-terminal protein domains. Accordingly, one thermal transition is observed for GyrA59.The melting profile of the GyrA subunit is dramatically affected by monovalent and divalent metal ions, both transitions being shifted to lower temperature upon increasing salt concentration. This effect is much more pronounced with divalent ions (Mg(2+)) and cannot be accounted for by changes in ionic strength only. The presence of ciprofloxacin shifts the melting transitions of the wild-type subunit to higher temperatures when physiological concentrations of Mg(2+) are present. In contrast, both the mutant protein and the 59 kDa fragment do not show evidence for quinolone-driven changes. These data suggest that ciprofloxacin binds to the wild-type subunit in an interaction that involves Ser83 of GyrA and that both C and N-terminal domains may be required for effective drug-protein interactions. The bell-shaped dependence of the binding process upon Mg(2+) concentration, with a maximum centred at 3-4 mM [Mg(2+)], is consistent with a metal-ion mediated GyrA-quinolone-interaction. Affinity chromatography data fully support these findings and additionally confirm the requirement for a free carboxylate to elicit binding of the quinolone to GyrA.We infer that the Mg(2+)-GyrA interaction at physiological metal ion concentration could bear biological relevance, conferring more conformational flexibility to the active enzyme. The results obtained in the presence of ciprofloxacin additionally suggest that the Mg(2+)-mediated quinolone binding to the enzyme might be involved in the mechanism of action of this family of drugs.  相似文献   

16.
The predicted active site of chorismate mutase of baker's yeast Saccharomyces cerevisiae has been studied by continuum electrostatics, molecular surface/volume calculations, and molecular modeling. Our study shows that despite being subject to an allosteric transition, the enzyme's active-site pocket neither decreased in volume nor deformed significantly in shape between the active R state and the inactive T state. We find that the polar atmosphere in the pocket is responsible for the enzyme's affinity. A single amino acid, Glu23, can adequately account for the atmospheric variation. This residue swings into the active-site pocket from the R state to the T state. In the R state, Glu23 on helix H2 doubly pairs with Arg204 and Lys208 of H11, which is packed against H2. In the T state, a slide occurs between H11 and H2 such that Glu23 can no longer interact with Lys208 and competes with Asp24 for interacting with Arg204. Consequently, Glu23 is found in the T state to couple with Arg157, an active-site residue critical to substrate binding. The tandem sliding of H11 in both monomers profoundly changes the interactions in the dimer interface. The loop between H11 and H12 demonstrates the largest conformational change. Hence, we establish a connection between the allosteric transition and the activity of the enzyme. The conformational change in the transition is suggested to propagate into the active-site pocket via a series of polar interactions that result in polarity reversal in the active-site pocket, which regulates the enzyme's activity. Proteins 31:445–452, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

17.
Fructose-1,6-bisphosphatase requires a divalent metal cation for catalysis, Mg(2+) being its most studied activator. Phosphatase activity increases sigmoidally with the concentration of Mg(2+), but the mechanistic basis for such cooperativity is unknown. Bound magnesium cations can interact within a single subunit or between different subunits of the enzyme tetramer. Mutations of Asp(118), Asp(121), or Glu(97) to alanine inactivate the recombinant porcine enzyme. These residues bind directly to magnesium cations at the active site. Three different hybrid tetramers of fructose-1,6-bisphosphatase, composed of one wild-type subunit and three subunits bearing one of the mutations above, exhibit kinetic parameters (K(m) for fructose-1,6-bisphosphate, 1.1-1.8 microm; K(a) for Mg(2+), 0.34-0.76 mm; K(i) for fructose-2,6-bisphosphate, 0.11-0.61 microm; and IC(50) for AMP, 3.8-7.4 microm) nearly identical to those of the wild-type enzyme. Notwithstanding these similarities, the k(cat) parameter for each hybrid tetramer is approximately one-fourth of that for the wild-type enzyme. Evidently, each subunit in the wild-type tetramer can independently achieve maximum velocity when activated by Mg(2+). Moreover, the activities of the three hybrid tetramers vary sigmoidally with the concentration of Mg(2+) (Hill coefficients of approximately 2). The findings above are fully consistent with a mechanism of cooperativity that arises from within a single subunit of fructose-1,6-bisphosphatase.  相似文献   

18.
During the reaction catalyzed by enolase, a mobile loop, residues 36-45, closes over the active site. In order to probe the role of this loop movement in catalysis, the glycines at positions 37 and 41 of rabbit muscle enolase (beta beta) have been mutated to alanines. The mutant forms-G37A and G41A-of enolase are both active, but have altered selectivity for divalent cations. G37A, when assayed with Mg(2+), has 12% the activity of the wild type. However, it is twice as active as wild type when assayed with Mn(2+), Zn(2+), or Co(2+). G41A has 4% the activity of the wild type with Mg(2+), is more active than wild type with Co(2+), and slightly less active than wild type with Mn(2+) and Zn(2+). The kinetic isotope effect for both mutants is greater than that of the wild type with all 4 divalent cations. These results indicate that the flexibility of this loop has subtle effects on catalytic activity.  相似文献   

19.
M D Sam  J J Perona 《Biochemistry》1999,38(20):6576-6586
The rate constant for the phosphoryl transfer step in site-specific DNA cleavage by EcoRV endonuclease has been determined as a function of pH and identity of the required divalent metal ion cofactor, for both wild-type and T93A mutant enzymes. These measurements show bell-shaped pH-rate curves for each enzyme in the presence of Mg2+ as a cofactor, indicating general base catalysis for the nucleophilic attack of hydroxide ion on the scissile phosphate, and general acid catalysis for protonation of the leaving 3'-O anion. The kinetic data support a model for phosphoryl transfer based on wild-type and T93A cocrystal structures, in which the ionizations of two distinct metal-ligated waters respectively generate the attacking hydroxide ion and the proton for donation to the leaving group. The model concurs with recent observations of two metal ions bound in the active sites of the type II restriction endonucleases BamHI and BglI, suggesting the possibility of a similar catalytic mechanism functioning in many or all members of this enzyme family.  相似文献   

20.
P-type ATPases such as the Na+,K+-ATPase (sodium pump) hydrolyze ATP to pump ions through biological membranes against their electrochemical gradients. The mechanisms that couple ATP hydrolysis to the vectorial ion transport are not yet understood, but unveiling structures that participate in ATP binding and in the formation of the ionophore might help to gain insight into this process. Looking at the alpha- and beta-phosphates of ATP as a pyrophosphate molecule, we found that peptides highly conserved among all soluble inorganic pyrophosphatases are also present in ion-transporting ATPases. Included therein are Glu48 and Lys56 of the Saccharomyces cerevisiae pyrophosphatase (SCE1-PPase) that are essential for the activity of this enzyme and have been shown in crystallographic analysis to interact with phosphate molecules. To test the hypothesis that equivalent amino acids are also essential for the activity of ion-transporting ATPases, Glu472 and Lys480 of the sodium pump alpha 1 subunit corresponding to Glu48 and Lys56 of SCE1-PPase were mutated to various amino acids. Mutants of the sodium pump alpha1 subunit were expressed in yeast and analyzed for their ATPase activity and their ability to bind ouabain in the presence of either ATP, Mg2+, and Na+ or phosphate and Mg2+. All four mutants investigated, Glu472Ala, Glu472Asp, Lys480Ala, and Lys480Arg, display only a fraction of the ATPase activity obtained with the wild-type enzyme. The same applies with respect to their ability to bind ouabain, where maximum ouabain binding to the mutants accounts for only about 10% of the binding obtained with the wild-type enzyme. On the basis of our results, we conclude that Glu472 and Lys480 are essential for the activity of the sodium pump. Their function is probably to arrest the alpha- and beta-phosphate groups of ATP in a proper position prior to hydrolysis of the gamma-phosphate group. The identification of these amino acids as essential components of the ATP-recognizing mechanism of the pump has resulted in a testable hypothesis for the initial interactions of the sodium pump, and possibly of other P-type ATPases, with ATP.  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