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.
The hammerhead cleavage reaction in monovalent cations   总被引:10,自引:3,他引:7       下载免费PDF全文
Recently, Murray et al. (Chem Biol, 1998, 5:587-595) found that the hammerhead ribozyme does not require divalent metal ions for activity if incubated in high (> or =1 M) concentrations of monovalent ions. We further characterized the hammerhead cleavage reaction in the absence of divalent metal. The hammerhead is active in a wide range of monovalent ions, and the rate enhancement in 4 M Li+ is only 20-fold less than that in 10 mM Mg2+. Among the Group I monovalent metals, rate correlates in a log-linear manner with ionic radius. The pH dependence of the reaction is similar in 10 mM Mg2+, 4 M Li+, and 4 M Na+. The exchange-inert metal complex Co(NH3)3+ also supports substantial hammerhead activity. These results suggest that a metal ion does not act as a base in the reaction, and that the effects of different metal ions on hammerhead cleavage rates primarily reflect structural contributions to catalysis.  相似文献   

2.
The apparent specificity of activation of lysine-sensitive aspartokinase (E.C.2.7.2.4) from E. coli by monovalent cations differs depending on the assay used and on the Mg2+ concentration. Activity is nearly absolutely dependent on and is highly specific for a monovalent cation in the aspartate semialdehyde dehydrogenase coupled assay or the adenosine triphosphate-adenosine diphosphate exchange assay. Little specificity for monovalent cations is observed using the aspartyl hydroxamate assay. Activation and specificity are also altered by Mg2+ concentrations at a constant 5 mM nucleotide concentration. At a low (1.25 or 1.6 mM)Mg2+ concentration, monovalent cation activation and specificity are nearly absolute. Less dependence on monovalent cations and less specificity are observed at a higher Mg2+ concentration (6 mM). Li+ inhibits aspartokinase competitively with respect to either K+ or NH4+. Monovalent cations are also thermoprotective and differential thermal inactivation experiments at 56 degrees C reveal that NH4+ and K+, either of which will produce maximum catalytic activity, interact differently with aspartokinase. K+ interacts with positive cooperativity, whereas NH4+ does not. K+, NH4+, and Na+ are about equally effective in enhancing the dissociation of the aspartokinase-aspartylphosphate complex. Li+ is less effective.  相似文献   

3.
The nitrogenase enzyme complex of Methanosarcina barkeri 227 was found to be more sensitive to NaCl than previously studied molybdenum nitrogenases are, with total inhibition of activity occurring at 190 mM NaCl, compared with >600 mM NaCl for Azotobacter vinelandii and Clostridium pasteurianum nitrogenases. Na+ and K+ had equivalent effects, whereas Mg2+ was more inhibitory than either monovalent cation, even on a per-charge basis. The anion Cl- was more inhibitory than acetate was. Because M. barkeri 227 is a facultative halophile, we examined the effects of external salt on growth and diazotrophy and found that inhibition of growth was not greater with N2 than with NH4+. Cells grown with N2 and cells grown with NH4+ produced equal concentrations of alpha-glutamate at low salt concentrations and equal concentrations of Nepsilon-acetyl-beta-lysine at NaCl concentrations greater than 500 mM. Despite the high energetic cost of fixing nitrogen for these osmolytes, we obtained no evidence that there is a shift towards nonnitrogenous osmolytes during diazotrophic growth. In vitro nitrogenase enzyme assays showed that at a low concentration (approximately 100 mM) potassium glutamate enhanced activity but at higher concentrations this compound inhibited activity; 50% inhibition occurred at a potassium glutamate concentration of approximately 400 mM.  相似文献   

4.
Escherichia coli ribonuclease III, purified to homogeneity from an overexpressing bacterial strain, exhibits a high catalytic efficiency and thermostable processing activity in vitro. The RNase III-catalyzed cleavage of a 47 nucleotide substrate (R1.1 RNA), based on the bacteriophage T7 R1.1 processing signal, follows substrate saturation kinetics, with a Km of 0.26 microM, and kcat of 7.7 min.-1 (37 degrees C, in buffer containing 250 mM potassium glutamate and 10 mM MgCl2). Mn2+ and Co2+ can support the enzymatic cleavage of the R1.1 RNA canonical site, and both metal ions exhibit concentration dependences similar to that of Mg2+. Mn2+ and Co2+ in addition promote enzymatic cleavage of a secondary site in R1.1 RNA, which is proposed to result from the altered hydrolytic activity of the metalloenzyme (RNase III 'star' activity), exhibiting a broadened cleavage specificity. Neither Ca2+ nor Zn2+ support RNase III processing, and Zn2+ moreover inhibits the Mg(2+)-dependent enzymatic reaction without blocking substrate binding. RNase III does not require monovalent salt for processing activity; however, the in vitro reactivity pattern is influenced by the monovalent salt concentration, as well as type of anion. First, R1.1 RNA secondary site cleavage increases as the salt concentration is lowered, perhaps reflecting enhanced enzyme binding to substrate. Second, the substitution of glutamate anion for chloride anion extends the salt concentration range within which efficient processing occurs. Third, fluoride anion inhibits RNase III-catalyzed cleavage, by a mechanism which does not involve inhibition of substrate binding.  相似文献   

5.
M R Ehlers  R E Kirsch 《Biochemistry》1988,27(15):5538-5544
The catalysis of the hydrolysis of angiotensin I, an important natural substrate, by human angiotensin-converting enzyme (ACE) was examined in detail as a function of chloride and hydrogen ion concentration. Chloride was found to be a nonessential activator over the pH range 5.0-10.0, with the chloride dependence increasing with increasing pH: the velocity enhancement at optimal [Cl-] increased from 1.6- to 42-fold; the chloride optimum and Ka' increased from 20 to 520 mM and from 0.22 to 120 mM, respectively, and activity in the absence of chloride decreased from 60.9 to 2.4% (relative to maximal activation). Kinetic analyses at pH 6.0, 7.5, and 9.0 confirmed the nonessential activator mechanism. At all pH values tested chloride was found to be inhibitory (relative to maximal activation) at supraoptimal chloride levels. Depending on the [Cl-] range, both apparent uncompetitive and competitive modes were demonstrated. From pH 6.0 to 9.0 Kis varied between 110 and 1140 mM (apparent). In all cases Ki' much greater than Ka'. We suggest that at high [Cl-] chloride binds to low-affinity inhibitory sites on the free enzyme and on the ES and EP complexes. The pH-rate profile demonstrated a chloride-dependent alkaline shift, with the pH optimum increasing from 7.1 at zero chloride to 7.6 at 400 mM NaCl. At [S] much greater than Km a plot of log nu vs pH revealed pKs of 5.9 and 9.4 in the ES complex in the absence of chloride, while at maximally activating [Cl-] only one ionization at pK = 6.3 was observed.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Fructose-1,6-bisphosphatase requires divalent cations (Mg2+, Mn2+, or Zn2+) for catalysis, but a diverse set of monovalent cations (K+, Tl+, Rb+, or NH(4)(+)) will further enhance enzyme activity. Here, the interaction of Tl+ with fructose-1,6-bisphosphatase is explored under conditions that support catalysis. On the basis of initial velocity kinetics, Tl+ enhances catalysis by 20% with a K(a) of 1.3 mm and a Hill coefficient near unity. Crystal structures of enzyme complexes with Mg2+, Tl+, and reaction products, in which the concentration of Tl+ is 1 mm or less, reveal Mg2+ at metal sites 1, 2, and 3 of the active site, but little or no bound Tl+. Intermediate concentrations of Tl+ (5-20 mm) displace Mg2+ from site 3 and the 1-OH group of fructose 6-phosphate from in-line geometry with respect to bound orthophosphate. Loop 52-72 appears in a new conformational state, differing from its engaged conformation by disorder in residues 61-69. Tl+ does not bind to metal sites 1 or 2 in the presence of Mg2+, but does bind to four other sites with partial occupancy. Two of four Tl+ sites probably represent alternative binding sites for the site 3 catalytic Mg2+, whereas the other sites could play roles in monovalent cation activation.  相似文献   

7.
8.
The acidic residues Asp-111, Asp-113, and Glu-115 of Escherichia coli DNA topoisomerase I are located near the active site Tyr-319 and are conserved in type IA topoisomerase sequences with counterparts in type IIA DNA topoisomerases. Their exact functional roles in catalysis have not been clearly defined. Mutant enzymes with two or more of these residues converted to alanines were found to have >90% loss of activity in the relaxation assay with 6 mM Mg(II) present. Mg(II) concentrations (15-20 mM) inhibitory for the wild type enzyme are needed by these double mutants for maximal relaxation activity. The triple mutant D111A/D113A/E115A had no detectable relaxation activity. Mg(II) binding to wild type enzyme resulted in an altered conformation detectable by Glu-C proteolytic digestion. This conformational change was not observed for the triple mutant or for the double mutant D111A/D113A. Direct measurement of Mg(II) bound showed the loss of 1-2 Mg(II) ions for each enzyme molecule due to the mutations. These results demonstrate a functional role for these acidic residues in the binding of Mg(II) to induce the conformational change required for the relaxation of supercoiled DNA by the enzyme.  相似文献   

9.
Accurate predictions of DNA stability in physiological and enzyme buffers are important for the design of many biological and biochemical assays. We therefore investigated the effects of magnesium, potassium, sodium, Tris ions, and deoxynucleoside triphosphates on melting profiles of duplex DNA oligomers and collected large melting data sets. An empirical correction function was developed that predicts melting temperatures, transition enthalpies, entropies, and free energies in buffers containing magnesium and monovalent cations. The new correction function significantly improves the accuracy of predictions and accounts for ion concentration, G-C base pair content, and length of the oligonucleotides. The competitive effects of potassium and magnesium ions were characterized. If the concentration ratio of [Mg (2+)] (0.5)/[Mon (+)] is less than 0.22 M (-1/2), monovalent ions (K (+), Na (+)) are dominant. Effects of magnesium ions dominate and determine duplex stability at higher ratios. Typical reaction conditions for PCR and DNA sequencing (1.5-5 mM magnesium and 20-100 mM monovalent cations) fall within this range. Conditions were identified where monovalent and divalent cations compete and their stability effects are more complex. When duplexes denature, some of the Mg (2+) ions associated with the DNA are released. The number of released magnesium ions per phosphate charge is sequence dependent and decreases surprisingly with increasing oligonucleotide length.  相似文献   

10.
Pyruvate phosphate dikinase (PPDK) catalyzes the interconversion of ATP, P(i), and pyruvate with AMP, PP(i), and phosphoenolpyruvate (PEP) in three partial reactions as follows: 1) E-His + ATP --> E-His-PP.AMP; 2) E-His-PP.AMP + P(i) --> E-His-P.AMP.PP(i); and 3) E-His-P + pyruvate --> E.PEP using His-455 as the carrier of the transferred phosphoryl groups. The crystal structure of the Clostridium symbiosum PPDK (in the unbound state) reveals a three-domain structure consisting of consecutive N-terminal, central His-455, and C-terminal domains. The N-terminal and central His-455 domains catalyze partial reactions 1 and 2, whereas the C-terminal and central His-455 domains catalyze partial reaction 3. Attempts to obtain a crystal structure of the enzyme with substrate ligands bound at the nucleotide binding domain have been unsuccessful. The object of the present study is to demonstrate Mg(II) activation of catalysis at the ATP/P(i) active site, to identify the residues at the ATP/P(i) active site that contribute to catalysis, and to identify roles for these residues based on their positions within the active site scaffold. First, Mg(II) activation studies of catalysis of E + ATP + P(i) --> E-P + AMP + PP(i) partial reaction were carried out using a truncation mutant (Tem533) in which the C-terminal domain is absent. The kinetics show that a minimum of 2 Mg(II) per active site is required for the reaction. The active site residues used for substrate/cofactor binding/activation were identified by site-directed mutagenesis. Lys-22, Arg-92, Asp-321, Glu-323, and Gln-335 mutants were found to be inactive; Arg-337, Glu-279, Asp-280, and Arg-135 mutants were partially active; and Thr-253 and Gln-240 mutants were almost fully active. The participation of the nucleotide ribose 2'-OH and alpha-P in enzyme binding is indicated by the loss of productive binding seen with substrate analogs modified at these positions. The ATP, P(i), and Mg(II) ions were docked into the PPDK N-terminal domain crevice, in an orientation consistent with substrate/cofactor binding modes observed for other members of the ATP-Grasp fold enzyme superfamily and consistent with the structure-function data. On the basis of this docking model, the ATP polyphosphate moiety is oriented/activated for pyrophosphoryl transfer through interaction with Lys-22 (gamma-P), Arg-92 (alpha-P), and the Gly-101 to Met-103 loop (gamma-P) as well as with the Mg(II) cofactors. The P(i) is oriented/activated for partial reaction 2 through interaction with Arg-337 and a Mg(II) cofactor. The Mg(II) ions are bound through interaction with Asp-321, Glu-323, and Gln-335 and substrate. Residues Glu-279, Asp-280, and Arg-135 are suggested to function in the closure of an active site loop, over the nucleotide ribose-binding site.  相似文献   

11.
Protein farnesyltransferase (FTase) catalyzes the addition of a farnesyl chain onto the sulfur of a C-terminal cysteine of a protein substrate. Magnesium ions enhance farnesylation catalyzed by FTase by several hundred-fold, with a KMg value of 4 mM. The magnesium ion is proposed to coordinate the diphosphate leaving group of farnesyldiphosphate (FPP) to stabilize the developing charge in the farnesylation transition state. Here we further investigate the magnesium binding site using mutagenesis and biochemical studies. Free FPP binds Mg2+ with a Kd of 120 microM. The 10-fold weaker affinity for Mg2+ observed for the FTase.FPP.peptide ternary complex is probably caused by the positive charges in the diphosphate binding pocket of FTase. Furthermore, mutation of aspartate beta 352 to alanine (D beta 352A) or lysine (D beta 352K) in FTase drastically alters the Mg2+ dependence of FTase catalysis without dramatically affecting the rate constant of farnesylation minus magnesium or the binding affinity of either substrate. In D beta 352A FTase, the KMg increases 28-fold to 110 +/- 30 mM, and the farnesylation rate constant at saturating Mg2+ decreases 27-fold to 0.30 +/- 0.05 s-1. Substitution of a lysine for Asp-beta 352 removes the magnesium activation of farnesylation catalyzed by FTase but does not significantly enhance the rate constant for farnesylation in the absence of Mg2+. In wild type FTase, Mg2+ can be replaced by Mn2+ with a 2-fold lower KMn (2 mM). These results suggest both that Mg2+ coordinates the side chain carboxylate of Asp-beta 352 and that the role of magnesium in the reaction includes positioning the FPP prior to catalysis.  相似文献   

12.
The aggregation equilibria of Escherichia coli RNA polymerase core and holoenzyme have been studied by velocity sedimentation as a function of [NaCl] both in the presence and in the absence of MgCl2. Effects of other anions (F- and I-), pH, and temperature have also been examined. Diffusion coefficients obtained by quasi-elastic light scattering (QLS) at high and low salt concentrations were used in conjunction with sedimentation coefficients under these conditions to obtain molecular weights of the protomer and aggregates of the core enzyme. At low salt concentration, core aggregates to a tetramer in the absence of MgCl2 and to an octamer in the presence of MgCl2. Some ambiguity exists in the interpretation of the sedimentation and QLS data for holoenzyme. The sedimentation results are consistent with the formation of dimers at low salt, both in the presence and in the absence of MgCl2. In all cases, equilibrium constants were calculated assuming a simple monomer--j-mer stoichiometry. These equilibrium constants are extremely sensitive functions of the concentration and type of monovalent anion. In Cl-, aggregation of both core and holoenzyme begins abruptly when the salt concentration is reduced below approximately 0.2 M (at a protein concentration of approximately 0.30 mg/mL); for core, substitution of I- for Cl- suppresses aggregation while F- enhances aggregation at a fixed anion concentration. No specific effect of monovalent cations (Na+, NH4+) is observed; Mg2+ has no effect on holoenzyme dimerization and has little effect on the salt range of core aggregation, though the stoichiometries of the core aggregates in the presence and absence of Mg2+ differ. Anion effects on these equilibria were modeled by assuming that a class of anion-binding sites on the protomer is not present in the aggregate, so that anion release accompanies aggregation. Analytical expressions for several models of the effect of anions on the aggregation equilibria were derived by using the method of binding polynomials. The salt dependence of the aggregation equilibria in the absence of Mg2+ appears inconsistent with a model in which the anion-binding sites on the protomer are independent (noncooperative), but it is well described by a model in which anion binding to the protomers occurs in a completely cooperative manner. The molecular basis of this apparent cooperative effect of anions on the aggregation equilibria is proposed to be an allosteric effect of anions on conformational equilibria of the protomers of core polymerase and the holoenzyme. Implications of such a salt-dependent conformational transition for the DNA-binding interactions of the enzyme are considered.  相似文献   

13.
Qian X  He Y  Luo Y 《Biochemistry》2007,46(20):5855-5863
RecA-like strand exchange proteins, which include closely related archaeal Rad51/RadA and eukaryal Rad51 and DMC1, play a key role in DNA repair by forming helical nucleoprotein filaments which promote a hallmark strand exchange reaction between homologous DNA substrates. Our recent crystallographic studies on a RadA recombinase from Methanococcus voltae (MvRadA) have unexpectedly revealed a secondary magnesium at the subunit interface approximately 11 A from the primary one coordinated by ATP and the canonical P-loop. The DNA-dependent ATPase activity of MvRadA appears to be dependent on the concentration of free Mg2+, while the strand exchange activity does not. We also made site-directed mutagenesis at the Mg2+-liganding residue Asp-246. The mutant proteins exhibited approximately 20-fold reduced ATPase activity but normal strand exchange activity. Structurally, the main chain carbonyl of the conserved catalytic residue Glu-151 is hydrogen bonded with one of the magnesium-liganding water molecules. Changes in the secondary magnesium site may therefore induce conformational changes around this catalytic glutamate and affect the ATPase activity without significantly altering the stability of the extended recombinase filament. Asp-246 is somewhat conserved among archaeal and eukaryal homologues, implying some homologues may share this allosteric site for ATPase function.  相似文献   

14.
The physiological concentration of free magnesium in Escherichia coli cells is about 1 mM, and there is almost no chloride in the cell. When the aminoacylation of tRNA by tyrosyl-tRNA synthetase was assayed at 1 mM free Mg2+, chloride (and sulphate) ions inhibited the reaction but acetate at the same concentration (< 200 mM) was not inhibitory. When the magnesium concentration was increased to 10 mM there was almost no chloride inhibition any more. Chloride strengthened the PPi inhibition, the Ki(app)(PPi) values at 1 mM free Mg2+ were 140, 120, and 56 microM at 0, 50 and 150 mM KCl, respectively. Chloride weakened the AMP inhibition, the corresponding values for Ki(app)(AMP) were 0.35, 0.5, and 0.9 mM. The value of Km(app)(tRNA(Tyr)) was clearly increased by chloride, being 22, 37, 93, and 240 nM at 0, 50, 100, and 150 mM KCl, respectively. Best-fit analyses of the PPi inhibition, AMP inhibition and Km(app)(tRNA) assays were accomplished using total rate equations. The analysis showed that the only kinetic events which are obligatory to explain the chloride effects are a weakened binding of Mg2+ to the tRNA before the transfer reaction and a weakened binding of Mg2+ to the Tyr-tRNA-enzyme complex after the transfer reaction. The dissociation constants for the former were 0.11, 0.3, and 2.8 mM and for the latter 0.6, 2.5, and 13 mM at 0, 50 and 150 mM KCl, respectively. Mg2+ is required for the reactive conformation of tRNA in the transfer reaction but chloride weakens its formation. After the transfer reaction the dissociation of Mg2+ from the aa-tRNA-enzyme complex enhances the dissociation of the aa-tRNA from the enzyme. The kinetics and the chloride effect were similar in the tyrosyl-tRNA synthetases from both Bacillus stearothermophilus and E. coli.  相似文献   

15.
Properties and reaction mechanism of C4 leaf pyruvate,Pi dikinase   总被引:3,自引:0,他引:3  
The properties and reaction mechanism of maize leaf pyruvate,Pi dikinase are described. Km values were determined for the forward reaction substrates, pyruvate, ATP, and Pi, at pH 7.4 and 8.0 and for reverse reaction substrates at pH 7.4. Enzyme activity was almost totally dependent on added monovalent cations in both directions. NH+4 was most effective, with Ka values of about 0.38 mM for the forward reaction and 2 mM for the reverse reaction. K+ also completely activated the enzyme in the forward direction (Ka = 8 mM) but only partially activated in the reverse direction. Na+ had little effect on either reaction. The pH optimum for the forward reaction was about 8.2; the reverse reaction optimum was about 6.9. Maximum activity for the reverse direction was about twice the maximum forward direction rate. From data on the requirements for the ATP-AMP exchange reaction, on the mechanism of inhibition of the forward reaction by PEP, AMP, and PPi, and from the kinetics of the interaction of varying certain substrate pairs, it was concluded that the maize leaf pyruvate,Pi dikinase reaction proceeded by the two-step Bi Bi Uni Uni mechanism. This differs from the mechanism of catalysis by the bacterial enzyme.  相似文献   

16.
General properties of ouabain-sensitive K+ binding to purified Na+,K+-ATPase [EC 3.6.1.3] were studied by a centrifugation method with 42K+. 1) The affinity for K+ was constant at pH values higher than 6.4, and decreased at pH values lower than 6.4. 2) Mg2+ competitively inhibited the K+ binding. The dissociation constant (Kd) for Mg2+ of the enzyme was estimated to be about 1 mM, and the ratio of Kd for Mg2+ to Kd for K+ was 120 : 1. The order of inhibitory efficiency of divalent cations toward the K+ binding was Ba2+ congruent to Ca2+ greater than Zn2+ congruent to Mn2+ greater than Sr2+ greater than Co2+ greater than Ni2+ greater than Mg2+. 3) The order of displacement efficiency of monovalent cations toward the K+ binding in the presence or absence of Mg2+ was Tl+ greater than Rb+ greater than or equal to (K+) greater than NH4+ greater than or equal to Cs+ greater than Na+ greater than Li+. The inhibition patterns of Na+ and Li+ were different from those of other monovalent cations, which competitively inhibited the K+ binding. 4) The K+ binding was not influenced by different anions, such as Cl-, SO4(2-), NO3-, acetate, and glycylglycine, which were used for preparing imidazole buffers. 5) Gramicidin D and valinomycin did not affect the K+ binding, though the former (10 micrograms/ml) inhibited the Na+,K+-ATPase activity by about half. Among various inhibitors of the ATPase, 0.1 mM p-chloromercuribenzoate and 0.1 mM tri-n-butyltin chloride completely inhibited the K+ binding. Oligomycin (10 micrograms/ml) and 10 mM N-ethylmaleimide had no effect on the K+ binding. In the presence of Na+, however, oligomycin decreased the K+ binding by increasing the inhibitory effect of Na+, whether Mg2+ was present or not. 6) ATP, adenylylimido diphosphate and ADP each at 0.2 mM decreased the K+ binding to about one-fourth of the original level at 10 microM K+ without MgCl2 and at 60 microM K+ with 5 mM MgCl2. On the other hand, AMP, Pi, and p-nitrophenylphosphate each at 0.2 mM had little effect on the K+ binding.  相似文献   

17.
Plastidic pyruvate kinase (ATP: pyruvate phosphotransferase, EC 2.7.1.40) was purified to near homogeneity as judged by native PAGE with about 4% recovery from developing seeds of Brassica campestris using (NH4)2SO4 fractionation, DEAE-cellulose chromatography, gel filtration through Sepharose-CL-6B and affinity chromatography through reactive blue Sepharose-CL-6B. The purified enzyme having molecular mass of about 266 kDa was quite stable and showed a broad pH optimum between pH 6.8-7.8. Typical Michaelis-Menten kinetics was obtained for both the substrates with K(m) values of 0.13 and 0.14 mM for PEP and ADP, respectively. The enzyme could also utilize CDP, GDP or UDP as alternative nucleotide to ADP, but with lower Vmax and higher K(m). The enzyme had an absolute requirement for a divalent and a monovalent cation for activity and was inhibited by oxalate, fumarate, citrate, isocitrate and ATP, and activated by AMP, aspartate, 3-PGA, tryptophan and inorganic phosphate. ATP inhibited the enzyme competitively with respect to PEP and non-competitively with respect to ADP. Similarly, oxalate inhibition was also of competitive type with respect to PEP and non-competitive with respect to ADP. This inhibition by either ATP or oxalate was not due to chelation of Mg2+, as the inhibition was not relieved on increasing Mg2+ concentration even upto 30 mM. Initial velocity and product inhibition studies demonstrated the reaction mechanism to be compulsory ordered type. The enzyme seems to be regulated synergistically by ATP and citrate.  相似文献   

18.
Dahlgren PR  Lyubchenko YL 《Biochemistry》2002,41(38):11372-11378
Atomic force microscopy (AFM) was applied to directly visualize the end-to-end DNA interaction mediated by magnesium cations. We took advantage of the APS-mica, allowing the preparation of samples in a broad range of monovalent and divalent cations to separate the effects of Mg(2+) and Na(+) cations on the interaction of restriction DNA fragments with cohesive end. The AFM data clearly show that DNA restriction fragments with cohesive ends form substantial amount of circles in the presence of Mg(2+) cations, suggesting that Mg(2+) cations stabilize the interaction of cohesive ends. This effect depends on the MgCl(2) concentration, so that the yield of circles approaches 18% in the presence of 50 mM MgCl(2). Furthermore, we demonstrate that this conferred cohesive end stability is specific for divalent cations, as substitution of MgCl(2) with NaCl leads to a near complete loss of cohesive end stability. We further demonstrate that cohesive end stabilization is achieved by substituting Mg(2+) with Ca(2+), Mn(2+), or Zn(2+). The data obtained suggest that the end stabilization mediated by divalent cations is primarily the result of inter-base interactions rather than bridging of phosphate moieties.  相似文献   

19.
The effects of monovalent ions on endogenous pyruvate dehydrogenase (PDH) kinase activity in purified bovine heart pyruvate dehydrogenase complex were investigated. Activity of PDH kinase was stimulated 1.9-, 1.95-, 1.65-, and 1.4-fold by 10 mM K+, Rb+, NH+4, and Cs+, respectively, whereas Na+ and Li+ had no effect on PDH kinase activity. The crystal radii of stimulatory ions were in the range of 1.33 to 1.69 A while the crystal radii of nonstimulatory ions were in the range of 0.6 to 0.94 A. Stimulation of PDH kinase by monovalent ions was not pH dependent. Protein dilution studies showed that monovalent ion stimulation was measurable within 10 s after protein addition to PDH kinase assays. Furthermore, stimulation occurred at all protein concentrations tested. At ATP concentrations from 12.5 to 25 microM, K+ and NH+4 stimulation was constant from 0 to 110 and 0 to 30 mM, respectively. At higher ATP concentrations, from 50 to 500 microM, K+ and NH+4 stimulation peaked at approximately 30 and 3 mM, respectively, and thereafter declined as the ion concentration increased. Maximal PDH kinase stimulation by K+ or NH+4 also declined as Na+ was increased from 0 to 120 mM, but at a fixed salt concentration of 120 mM, both K+ and NH+4 stimulated PDH kinase activity. Phosphopeptide analysis demonstrated that K+ and NH+4 stimulated phosphorylation at sites 1 and 2, but that site 3 phosphorylation was relatively constant under all conditions. Thiamin pyrophosphate and 5,5'-dithiobis-(2-nitrobenzoate) blocked monovalent ion stimulation half-maximally at 4 and 6 microM, respectively. However, neither thiamin pyrophosphate nor 5,5'-dithiobis-(2-nitrobenzoate) significantly inhibited PDH kinase activity in the absence of monovalent ions. The results indicate that heart PDH kinase stimulation by monovalent ions does not occur by changing the binding equilibrium between PDH and dihydrolipoyl transacetylase core. Instead, monovalent ions bind and exert their regulatory effects at or near the active site of PDH kinase.  相似文献   

20.
The objective of this study is to quantify the contributions of cations, anions and water to stability and specificity of the interaction of lac repressor (lac R) protein with the strong-binding symmetric lac operator (Osym) DNA site. To this end, binding constants Kobs and their power dependences on univalent salt (MX) concentration (SKobs = d log Kobs/d log[MX]) have been determined for the interactions of lac R with Osym operator and with non-operator DNA using filter binding and DNA cellulose chromatography, respectively. For both specific and non-specific binding of lac R, Kobs at fixed salt concentration [KX] increases when chloride (Cl-) is replaced by the physiological anion glutamate (Glu-). At 0.25 M-KX, the increase in Kobs for Osym is observed to be approximately 40-fold, whereas for non-operator DNA the increase in Kobs is estimated by extrapolation to be approximately 300-fold. For non-operator DNA, SKobsRD is independent of salt concentration within experimental uncertainty, and is similar in KCl (SKobs,RDKCl = -9.8(+/- 1.0) between 0.13 M and 0.18 M-KCl) and KGlu (SKobs,RDKGlu = -9.3(+/- 0.7) between 0.23 M and 0.36 M-KGlu). For Osym DNA, SKobsRO varies significantly with the nature of the anion, and, at least in KGlu appears to decrease in magnitude with increasing [KGlu]. Average magnitudes of SKobsRO are less than SKobsRD, and, for specific binding decrease in the order [SKobsRO,KCl[>[SKobsRO,KAc[>[SKobsRO,KGlu[ . Neither KobsRO nor SKobsRO is affected by the choice of univalent cation M+ (Na+, K+, NH4+, or mixtures thereof, all as the chloride salt), and SKobsRO is independent of [MCl] in the range examined (0.125 to 0.3 M). This behavior of SKobsRO is consistent with that expected for a binding process with a large contribution from the polyelectrolyte effect. However, the lack of an effect of the nature of the cation on the magnitude of KobsRO at a fixed [MX] is somewhat unexpected, in view of the order of preference of cations for the immediate vicinity of DNA (NH4+ > K+ > Na+) observed by 23Na nuclear magnetic resonance. For both specific and non-specific binding, the large stoichiometry of cation release from the DNA polyelectrolyte is the dominant contribution to SKobs. To interpret these data, we propose that Glu- is an inert anion, whereas Ac- and Cl- compete with DNA phosphate groups in binding to lac repressor. A thermodynamic estimate of the minimum stoichiometry of water release from lac repressor and Osym operator (210(+/- 30) H2O) is determined from analysis of the apparently significant reduction in [SKobsRO,KGlu[ with increasing [KGlu] in the range 0.25 to 0.9 M. According to this analysis, SKobs values of specific and non-specific binding in KGlu differ primarily because of the release of water in specific binding. In KAc and KCl, we deduce that anion competition affects Kobs and SKobs to an extent which differs for different anions and for the different binding modes.  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