首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 827 毫秒
1.
The parvalbumin metal ion-binding sites differ at the +z and -x residues: Whereas the CD site employs serine and glutamate (or aspartate), respectively, the EF site employs aspartate and glycine. Although frequently indistinguishable in Ca2+- and Mg2+-binding assays, the CD and EF sites nonetheless exhibit markedly different preferences for members of the lanthanide series [Williams et al. (1984) J. Am. Chem. Soc. 106, 5698-5702], underscoring an intrinsic nonequivalence. This nonequivalence reaches its pinnacle in the mammalian beta-parvalbumin (oncomodulin). Whereas the oncomodulin EF site exhibits the expected Ca2+/Mg2+ signature, the Ca2+ affinity of the CD site is severely attenuated. To obtain insight into the structural factors responsible for this reduction in binding affinity, oncomodulin variants were examined in which the CD and EF site ligand arrays had been exchanged. Our data suggest that binding affinity may be dictated either by ligand identity or by the binding site environment. For example, the Ca2+ affinity of the quasi-EF site resulting from the combined S55D and D59G mutations is substantially lower than that of the authentic EF site. This finding implies that other local environmental variables (e.g., binding loop flexibility, electrostatic potentials) within the CD binding site supersede the influence of ligand identity. However, the CD site ligand array does not acquire a high-affinity signature when imported into the EF site, as in the D94S/G98D variant. Instead, it retains its Ca2+-specific signature, implying that this constellation of ligands is less sensitive to placement within the protein molecule. The D59G and D94S single mutations substantially lower binding affinity, consistent with removal of a liganding carboxylate. By contrast, the S55D and G98D mutations substantially increase binding affinity, a finding at odds with corresponding data collected on model peptide systems. Significantly, the Ca2+ affinity of the oncomodulin CD site is increased by mutations that weaken binding at the EF site, indicating a negatively cooperative interaction between the two sites.  相似文献   

2.
Mapping of the conserved sequence regions in the restriction endonucleases MunI (C/AATTG) and EcoRI (G/AATTC) to the known X-ray structure of EcoRI allowed us to identify the sequence motif 82PDX14EXK as the putative catalytic/Mg2+ ion binding site of MunI [Siksnys, V., Zareckaja, N., Vaisvila, R., Timinskas, A., Stakenas, P., Butkus, V., & Janulaitis, A. Gene (1994) 142, 1-8]. Site-directed mutagenesis was then used to test whether amino acids P82, D83, E98, and K100 were important for the catalytic activity of MunI. Mutation P82A generated only a marginal effect on the cleavage properties of the enzyme. Investigation of the cleavage properties of the D83, E98, and K100 substitution mutants, however, in vivo and in vitro, revealed either an absence of catalytic activity or markedly reduced catalytic activity. Interestingly, the deleterious effect of the E98Q replacement in vitro was partially overcome by replacement of the metal cofactor used. Though the catalytic activity of the E98Q mutant was only 0.4% of WT under standard conditions (in the presence of Mg2+ ions), the mutant exhibited 40% of WT catalytic activity in buffer supplemented with Mn2+ ions. Further, the DNA binding properties of these substitution mutants were analyzed using the gel shift assay technique. In the absence of Mg2+ ions, WT MunI bound both cognate DNA and noncognate sequences with similar low affinities. The D83A and E98A mutants, in contrast, in the absence of Mg2+ ions, exhibited significant specificity of binding to cognate DNA, suggesting that the substitutions made can simulate the effect of the Mg2+ ion in conferring specificity to the MunI restriction enzyme.  相似文献   

3.
Nicotinamide nucleotide transhydrogenase catalyzes the reversible reduction of NADP+ by NADH and a concomitant proton translocation. It was demonstrated (Glavas, N.A. and Bragg, P.D. (1995) Biochim. Biophys. Acta 1231, 297-303) that the Escherichia coli transhydrogenase also catalyzed a reduction of the NAD-analogue 3-acetylpyridine-NAD+ (AcPyAD+) by NADH at low pH and in the absence of (added) NADP(H) and high salt concentrations The mechanism of this reaction has as yet not been explained. In the present study, the E. coli transhydrogenase was purified by affinity chromatography through the NADP(H)-site, rendering the pure enzyme free of NADP(H). Using this preparation it was confirmed that the enzyme readily catalyzes the above reaction. Inhibitors specific for the NADP(H)-site, e.g., palmitoyl-Coenzyme A and adenosine-2'-monophosphate-5'-diphosphoribose, strongly inhibited the reduction of AcPyAD+ by NADH, whereas an inhibitor of the NAD(H)-site, adenosine 5'-diphosphoribose, was less inhibitory. This suggests that a lack of metal ions or other ions at low pH induces an unspecific interaction of the NADP(H)-site with AcPyAD+ or NADH, presumably NADH, producing a cyclic reduction of AcPyAD+ by NADH via NAD(H) bound in the NADP(H) site. A stimulation of reduction of AcPyAD+ by NADPH by Mg2+ present during reconstitution of transhydrogenase in phospholipid vesicles was observed, but it is presently unclear whether this effect is related to that seen with the detergent-dispersed enzyme.  相似文献   

4.
The Na,K-ATPase activity of the sodium pump exhibits apparent multisite kinetics toward ATP, a feature that is inherent to the minimal enzyme unit, the alpha beta protomer. We have argued that this should arise from separate catalytic and noncatalytic sites on the alpha beta protomer as fluorescein isothiocyanate (FITC) blocks a high affinity ATP site on all alpha subunits and yet the modified Na, K-ATPase retains a low affinity response to nucleotides (Ward, D. G., and Cavieres, J. D. (1996) J. Biol. Chem. 271, 12317-12321). We now find that 2'(3')-O-(2,4,6-trinitrophenyl)8-azido-adenosine 5'-diphosphate (TNP-8N3-ADP), a high affinity photoactivatable analogue of ATP, can inhibit the K+-phosphatase activity of the FITC-modified enzyme during assays in dimmed light. The inhibition occurs with a Ki of 140 microM at 20 mM K+; it requires the adenine ring as 2'(3')-O-(2,4 6-trinitrophenyl) (TNP)-UDP or TNP-uridine are less potent and 2,4,6-trinitrobenzene-sulfonate is ineffective. Under irradiation with UV light, TNP-8N3-ADP inactivates the K+-phosphatase activity of the fluorescein-enzyme and also its phosphorylation by [32P]Pi. The photoinactivation process is stimulated by Na+ or Mg2+, and is inhibited by K+ or excess TNP-ADP. In the presence of 50 mM Na+ and 1 mM Mg2+, TNP-8N3-ADP photoinactivates with a K0.5 of 15 microM. Furthermore, TNP-8N3-ADP photoinactivates the FITC-modified, solubilized alpha beta protomers, even more effectively than the membrane-bound fluorescein-enzyme. These results strongly suggest that catalytic and allosteric ATP sites coexist on the alpha beta protomer of Na,K-ATPase.  相似文献   

5.
Expression of Na, K-ATPase in yeast allowed targeting of alpha beta-units with lethal substitutions at the phosphorylation site alpha 1 (D369N) beta 1 and alpha 1 (D369A) beta 1 at the cell surface at the same concentration of alpha-subunit and [3H] ouabain binding sites as for wild type Na, K-ATPase. Phosphorylation and reaction with vanadate were abolished, and the mutations had no Na, K-ATPase or K-phosphatase activity. Binding of [3H]-ATP at equilibrium revealed an intrinsic high affinity of the D369A mutation for ATP (KD = 2.8 nM) that was 39-fold higher than for wild type Na, K-ATPase (KD = 109 nM). The affinities for ADP were unaffected, indicating that the negative charge at residue 369 determines the contribution of the gamma-phosphate to the free energy of ATP binding. Analysis of the K(+)-ATP antagonism showed that the reduction of charge and hydrophobic substitution at Asp369 of the alpha-subunit caused a large shift in conformational equilibrium toward the E2-form. This was accompanied by a large increase in affinity for [3H] ouabain in Mg2+ medium with KD = 4.9 nM for D369A compared to KD = 51 nM for D369N and KD = 133 nM for wild type, and [3H] ouabain binding (KD = 153 nM) to D369A was detectable even in absence of Mg2+. In addition to its function as receptor of the gamma-phosphate of ATP, Asp369 has important short-range catalytic functions in modulating the affinity for ATP and long-range functions in governing the E1-E2 transitions which are coupled to reorientation of cation sites and changes in affinity for digitalis glycosides.  相似文献   

6.
A mutant of Lactobacillus casei dihydrofolate reductase, D26N, in which the active site aspartic acid residue has been replaced by asparagine by oligonucleotide-directed mutagenesis has been studied by NMR and optical spectroscopy and its kinetic behavior characterized in detail. On the basis of comparisons of a large number of chemical shifts and NOEs, it is clear that there are only very slight structural differences between the methotrexate complexes of the wild-type and mutant enzymes and that these are restricted to the immediate environment of the substitution. The data suggest a slight difference in orientation of the pteridine ring in the binding site in the mutant enzyme. Both NMR and UV spectroscopy show that methotrexate is protonated on N1 when bound to the wild-type enzyme but not when bound to the mutant. Binding constant measurements by fluorescence quenching and steady-state kinetic measurements of dihydrofolate (FH2) and folate reduction show that the substitution has little or no effect on substrate, coenzyme, and inhibitor binding (< 7-fold increase in Kd) and only a modest effect on kcat (up to a factor of 9 for FH2 and 25 for folate) and kcat/KM (up to a factor of 13 for FH2 and 14 for folate). Measurements of deuterium isotope effects and direct measurements of hydride ion transfer and product release by stopped-flow methods revealed that for the mutant enzyme hydride ion transfer is rate-limiting across the pH range 5-8. This allowed a direct comparison of the rate of hydride ion transfer in the wild-type and mutant enzymes; the asparagine substitution was found to decrease this rate by 62-fold at pH 5.5 and 9-fold at pH 7.5. This effect is much smaller than that seen for the corresponding mutant of Escherichia coli dihydrofolate reductase [Howell, E. E., Villafranca, J. E., Warren, M. S., Oatley, S. J., & Kraut, J. (1986) Science 231, 1123-1128], estimated as a 1000-fold decrease in the rate of hydride ion transfer. The change in pH dependence of kcat resulting from the substitution is consistent with, but does not prove, the idea that the group of pK 6.0 which must be protonated for hydride ion transfer to occur is Asp26. For folate reduction, the pH dependence of kcat is determined by two pKs, one of which, pK 5, disappears in the mutant enzyme, suggesting that it may correspond to ionization of Asp26.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
The role of Mg2+ in dTTP hydrolysis, dTTP binding, hexamer formation, and DNA binding was studied in bacteriophage T7 DNA helicase (4A' protein). The steady state kcat for the dTTPase activity was 200-300-fold lower in the absence of MgCl2, but the Km was only slightly affected. Direct dTTP binding experiments showed that the Kd of dTTP was unaffected, but the stoichiometry of dTTP binding was different in the absence of Mg2+. Two dTTPs were found to bind tightly in the absence of Mg2+ in contrast to three to four in the presence of Mg2+. In the presence of DNA there was little difference in the stoichiometry of dTTP binding to 4A'. These results indicate that Mg2+ is not necessary for dTTP binding, but Mg2+ is required for optimal hydrolysis of dTTP. Gel filtration of 4A' in the presence of dTTP without Mg2+ showed that Mg2+ was not necessary, and dTTP was sufficient for hexamer formation. The hexamers formed in the presence of dTTP without Mg2+ were capable of binding single-stranded DNA. However, the 4A' hexamers formed in the presence of dTDP with or without Mg2+ did not bind DNA, indicating that hexamer formation itself is not sufficient for DNA binding. The hexamers need to be in the correct conformation, in this case in the dTTP-bound state, to interact with the DNA. Thus, the gamma-phosphate of dTTP plays an important role in causing a conformational change in the protein that leads to stable interactions of 4A' with the DNA.  相似文献   

8.
The modes of Mg2+ binding to SMase from Bacillus cereus were studied on the basis of the changes in the tryptophyl fluorescence intensity. This enzyme was shown to possess at least two binding sites for Mg2+ with low and high affinities. The effects of Mg2+ binding on the enzymatic activity and structural stability of the enzyme molecule were also studied. The results indicated that the binding of Mg2+ to the low-affinity site was essential for the catalysis, but was independent of the substrate binding to the enzyme. It was also indicated that the alkaline denaturation of the enzyme was partly prevented by the Mg2+ binding, whereas no significant protective effect was observed against the denaturation by urea. The pH dependence of the kinetic parameters for the hydrolysis of micellar HNP and mixed micellar SM with Triton X-100 (1:10), catalyzed by SMase from B. cereus, was studied in the presence of a large amount of Mg2+ to saturate both the low- and high-affinity sites. The pH dependence curves of the logarithm of 1/Km for these two kinds of substrates were similar in shape to each other, and showed a single transition. On the other hand, the shapes of the pH dependence curves of the logarithm of kcat for these two kinds of substrates were different from each other. The pH dependence curve for micellar HNP showed three transitions and, counting from the acidic end of the pH region, the first and third transitions having tangent lines with slopes of +1 and -1, respectively. On the other hand, the curve for mixed micellar SM with Triton X-100 showed one large transition with a slope of +1 (the first transition) and a very small transition (the third transition). On the basis of the present results and the three-dimensional structure of bovine pancreatic DNase I, which has a primary structure similar to that of B. cereus SMase, we proposed a catalytic mechanism for B. cereus SMase based on general-base catalysis.  相似文献   

9.
The three-dimensional structures of E. coli inorganic pyrophosphatase (PPase) and its complexes with Mn2+ in a high affinity site and with Mg2+ in high and low affinity sites determined by authors in 1994-1996 at 1.9-2.2 A resolution are compared. Metal ion binding initiates the shifts of alpha-carbon atoms and of functional groups and rearrangement of non-covalent interaction system of hexameric enzyme molecule. As a result, the apoPPase with six equal subunits turns after Mg2+ binding into the structure with three types of subunits distinguished by structure and occupance of the low affinity Mg2+ site. Induced asymmetry reflects the subunit interactions and cooperativity between Mg2+ binding sites. These molecular rearrangements are structural basis to account for special features of the enzyme behavior and to propose one of the pathways for enzymatic activity regulation of constitutive PPases in vivo.  相似文献   

10.
Metal cofactors (Mg2+ and Mn2+) modulate both specific DNA binding and strand cleavage in the TaqI endonuclease (Cao, W., Mayer, A. N., and Barany, F. (1995) Biochemistry 34, 2276-2283). This work attempts to establish the structural basis of TaqI-DNA-metal2+ interactions using an affinity cleavage technique. The protein was cleaved by localized hydroxyl radicals generated by oxidizing Fe2+ within the metal binding sites. Cleavage fragments were separated by SDS-polyacrylamide gel electrophoresis, and cleavage sites were determined using micropeptide sequencing. Eleven amino acid residues in the vicinity of cleavage sites were selected for site-directed mutagenesis. The negative charge at Asp137 is essential for DNA cleavage but not required for sequence specific binding. Mutations at Asp142 abolish both specific binding and catalysis, except for D142E, which converts TaqI into a completely Mn2+-dependent endonuclease. The positive charge at Lys158 appears to be important for both specific binding and catalysis. Mutations at other sites affect binding and/or catalysis to different degrees, except Trp113 and Glu135, which appear to be nonessential for the TaqI enzyme activity. The critical residues for TaqI function are distinct from the PDX14-20(E/D)XK catalytic motif elucidated from other endonucleases.  相似文献   

11.
The antibiotic phosphomycin (1,2-epoxypropylphosphonic acid), an analog of phosphoenolpyruvate (PEP), behaved not as an inhibitor, but as an activator, of the enzyme phosphoenolpyruvate carboxylase (PEPC) from maize leaves. Multiple activation studies indicated that the analog binds to the Glc6P-allosteric site producing a more activated enzyme than Glc6P itself. Because of this, we used phosphomycin as a tool to further extend our understanding of the mechanisms of allosteric regulation of C4-PEPC. Initial velocity data from detailed kinetic studies, in which the concentrations of free and Mg-complexed PEP and phosphomycin were controlled, are consistent with: (1) the true activator is free phosphomycin, which competes with free PEP for the Glc6P-allosteric site; and (2) the Mg-phosphomycin complex caused inhibition by binding to the active site in competition with MgPEP. Therefore, although the Glc6P-allosteric site and the active site are able to bind the same ligands, they differ in the form of substrate and activator they bind. This important difference allows the full expression of the potential of activation and prevents inhibition by the activators, including the physiological ones, which are mostly uncomplexed at physiological free Mg2+ concentrations. At fixed low substrate concentrations, the saturation kinetics of the enzyme by phosphomycin showed positive cooperativity at pH 7.3 and 8.3, although at the latter pH, the kinetics of saturation by the substrate was hyperbolic. The cosolute glycerol greatly increased the affinity of the enzyme for phosphomycin and abolished the cooperativity in its binding, but did not eliminate the heterotropic effects of the activator. Therefore, the heterotropic and homotropic effects of the activator are not always coupled to the homotropic effects of the substrate, which argues against the two-state model previously proposed to explain the allosteric properties of maize-leaf PEPC.  相似文献   

12.
Cholinesterases exhibit functions apart from their esterase activity. We have demonstrated an aryl acylamidase and a zinc stimulated metallocarboxypeptidase activity in human serum butyrylcholinesterase. To establish the presence of zinc binding sites in the enzyme we examined the effect of metal chelators on its catalytic activities. The metal chelators 1,10-phenanthroline and N,N,N',N'-tetrakis (2-pyridyl methyl)ethylene diamine (TPEN) inhibited all the three catalytic activities in the enzyme. However, EDTA inhibited the peptidase activity exclusively without affecting the cholinesterase and aryl acylamidase activities. The catalytic activities were recovered upon removal of the chelator by Sephadex G-25 chromatography. Pre-treatment of the enzyme with any one of the three chelators resulted in the binding of the enzyme to a zinc-Sepharose column or to 65Zn2+. Histidine modification of the enzyme pretreated with chelators resulted in abolition of 65Zn2+ binding and zinc-Sepharose binding. Whereas the binding studies demonstrated removal of a metal from a Zn2+ binding site, attempts to remove the metal responsible for catalytic activity were unsuccessful. Atomic absorption spectroscopy indicated approximately 2.5 mol of zinc per mol of enzyme before treatment with EDTA and 1 mol zinc per mol enzyme after EDTA treatment. The results indicate that there are at least two metal binding sites on butyrycholinesterase. The presence of two HXXE...H sequences in butyrylcholinesterase supports these findings. Our studies implicate a zinc dependent metallocarboxypeptidase activity in the non-cholinergic functions of butyrylcholinesterase.  相似文献   

13.
The structures of Escherichia coli soluble inorganic pyrophosphatase (E-PPase) and Thermus thermophilus soluble inorganic pyrophosphatase (T-PPase) have been compared to find the basis for the superior thermostability of T-PPase. Both enzymes are D3 hexamers and crystallize in the same space group with very similar cell dimensions. Two rather small changes occur in the T-PPase monomer: a systematic removal of Ser residues and insertion of Arg residues, but only in the C-terminal part of the protein, and more long-range ion pairs from the C-terminal helix to the rest of the molecule. Apart from the first five residues, the three-dimensional structures of E-PPase and T-PPase monomers are very similar. The one striking difference, however, is in the oligomeric interactions. In comparison with an E-PPase monomer, each T-PPase monomer is skewed by about 1 A in the xy plane, is 0.3 A closer to the center of the hexamer in the z direction, and is rotated by approximately 7 degrees about its center of gravity. Consequently, there are a number of additional hydrogen bond and ionic interactions, many of which form an interlocking network that covers all of the oligomeric surfaces. The change can also be seen in local distortions of three small loops involved in the oligomeric interfaces. The complex rigid-body motion has the effect that the hexamer is more tightly packed in T-PPase: the amount of surface area buried upon oligomerization increases by 16%. The change is sufficiently large to account for all of the increased thermostability of T-PPase over E-PPase and further supports the idea that bacterial PPases, most active as hexamers or tetramers, achieve a large measure of their stabilization through oligomerization. Rigid-body motions of entire monomers to produce tighter oligomers may be yet another way in which proteins can be made thermophilic.  相似文献   

14.
Heats of CO and IHP binding to hemoglobin A have been determined under a variety of buffer and pH conditions. From these data heats of ion binding linked to hemoglobin oxygenation have been estimated. For IHP binding to deoxyhemoglobin the buffer-corrected enthalpies are surprisingly large, reaching -25 kcal/mol of IHP at pH 7.4. These values correspond to approximately -11 kcal/mol of proton absorbed upon IHP binding and may rise largely from the protonation of hitidine and NH2-terminal groups in the binding site (Arnone, A., and Perutz, M.F. (1974) Nature 249, 34-36). The decreased magnitude of delta HIHP observed at low pH parallels the decreased proton uptake at low pH. In 0.1 M chloride (pH 7.4) the reaction Hb(aq) + IHP leads to Hb x IHP(aq) has a standard free energy change (Edalji, R., Benesch, R.E., and Benesch, R. (1976) J. Biol. Chem. 251, 7720-7721) of -10 kcal and an enthalpy change of -25 kcal. Therefore, enthalpic forces provide the dominant driving force of this process. The origin of these large negative enthalpy changes is attributed to the exothermic protonation of protein basic groups induced by the proximity of phosphate negative charges. The importance of protonation in the binding of organic phosphates to hemoglobin may well extend to the specific binding of other phosphate substrates to enzyme reaction sites.  相似文献   

15.
16.
The hydrolytic properties of the alpha3beta3gamma and mutant alpha3(betaY341W)3gamma subcomplexes of the TF1-ATPase have been compared. ATPase activity of the mutant is less sensitive to turnover-dependent inhibition by azide, less suppressed by increasing concentrations of Mg2+ during assay, and less stimulated by lauryl dimethylamine oxide (LDAO). Therefore, it has much lower propensity than wild-type to entrap inhibitory MgADP in a catalytic site during turnover. The fluorescence of the introduced tryptophans in the alpha3(betaY341W)3gamma subcomplex is completely quenched when catalytic sites are saturated with ATP or ADP with or without Mg2+ present. As reported for the betaY331W mutant of Escherichia coli F1 (Weber, J., Wilke-Mounts, S., Lee, R. S.-F., Grell, E., Senior, A. E. (1993) J. Biol. Chem. 268, 20126-20133), this provides a direct probe of nucleotide binding to catalytic sites. Addition of stoichiometric MgATP to the mutant subcomplex quenched one-third the tryptophan fluorescence which did not recover after 60 min. This was caused by entrapment of MgADP in a single catalytic site. Titration of catalytic sites of the alpha3(betaY341W)3gamma subcomplex with MgADP or MgATP revealed Kd's of < 50 nM, about 0.25 microM and about 35 microM. Titrations were not affected by azide, whereas LDAO lowered the affinities of catalytic sites 2 and 3 for MgADP by 5-fold and 2-fold, respectively. During titration with MgATP, LDAO slightly lowered affinity at ATP concentrations below 30 microM and had no effect at ATP concentrations above 30 microM. Maximal velocity was attained when the third catalytic site was titrated with MgATP in the presence or absence of LDAO. The same Kd's for binding MgATP to the (alphaA396C)3beta3(gammaA22C) mutant were observed before and after inactivating it by cross-linking alpha to gamma. This implies that the different affinities of catalytic sites for MgATP do not represent negative cooperativity, but rather represent heterogeneous affinities of catalytic sites dictated by the position of the coiled-coil of the gamma subunit within the central cavity of the (alpha beta)3 hexamer.  相似文献   

17.
Chicken liver phosphoenolpyruvate carboxykinase (PEPCK) was rapidly inactivated by micromolar concentrations of ferrous sulfate in the presence of ascorbate at pH 7.4. Omitting ascorbate or replacing the Fe2+ with Mn2+ or Mg2+ gives no inactivation. Mn2+, Mg2+, or Co2+ at 100-fold molar excess over Fe2+ offered complete protection from Fe2+/ascorbate-induced inactivation. The substrates PEP and GTP, but not OAA, GDP, or CO2, offered full protection from inactivation. The addition of 5 mM EDTA stopped further inactivation of the enzyme. Thermodynamic studies indicate that the inactive enzyme no longer binds Mn2+ but still had high affinity for GTP indicating that the inactivation process was specific for the metal site. A decrease in cysteine content was observed over time following PEPCK treatment with Fe2+ and ascorbate. The apparent first-order rate constant for free sulfhydryl loss (0.085 +/- 0.005 min-1) is similar to the apparent first-order rate constant for inactivation (0.067 +/- 0.005 min-1). Amino acid composition analysis revealed that cysteic acid was generated upon Fe2+/ascorbate addition to PEPCK. Native chicken liver PEPCK has an Mr of 67 kDa. SDS-PAGE of the inactivated enzyme showed the presence of two new bands at 31.7 and 35.3 kDa indicating that PEPCK was specifically cleaved at a single site. The rate of cleavage was slower than the rate of inactivation and fully inactivated enzyme was only 50% cleaved. The Fe2+/ascorbate-catalyzed inactivation was not solely due to protein cleavage. The protein fragments generated by cleavage were separated by C4 reverse phase HPLC. The cleavage exposed a new N-terminus which was identified to be the 35.3 kDa C-terminal half of PEPCK. Sequencing of the fragments indicated that the site of cleavage was between Asp296 and Ile297. These results indicate that Asp296 is involved in metal chelation. This agrees with previous studies [Hlavaty, J. J., & Nowak, T. (1997) Biochemistry 36, 3389-3403] that suggested that Asp295 and Asp296 are involved in metal binding.  相似文献   

18.
3beta-(Iodoacetoxy)dehydroisoandrosterone (3beta-IDA), an analogue of the electrophilic substrate, Delta5-androstene-3,17-dione, as well as an analogue of several other steroid inhibitors of glutathione S-transferase, was tested as an affinity label of rat liver glutathione S-transferase, isozyme 1-1. A time-dependent loss of enzyme activity is observed upon incubation of 3beta-IDA with the enzyme. The rate of enzyme inactivation exhibits a nonlinear dependence on 3beta-IDA concentration, yielding an apparent Ki of 21 microM. Upon complete inactivation of the enzyme, a reagent incorporation of approximately 1 mol/mol of enzyme subunit or 2 mol/mol of enzyme dimer is observed. Protection against inactivation and incorporation is afforded by alkyl glutathione derivatives and nonsubstrate steroid ligands such as 17beta-estradiol-3,17-disulfate but, surprisingly, not by Delta5-androstene-3,17-dione or any other electrophilic substrate analogues tested. These results suggest that the site of reaction is within the nonsubstrate steroid binding site of the enzyme, which is distinguishable from the electrophilic substrate binding site, near the active site of the enzyme. Two cysteine residues, Cys17 and Cys111, are modified in nearly equal amounts, despite an average reagent incorporation of 1 mol/mol enzyme subunit. Isolation of enzyme subunits indicates the presence of unmodified, singly labeled, and doubly labeled subunits, consistent with mutually exclusive modification of cysteine residues across enzyme subunits; i.e., modification of Cys111 on subunit A prevents modification of Cys111 on subunit B and similarly for Cys17. Molecular modeling analysis suggests that Cys17 and Cys111 are located in the nonsubstrate steroid binding site, within the cleft between the subunits of the dimeric enzyme.  相似文献   

19.
The recent determination of the crystal structure of adenylyl cyclase has elucidated many structural features that determine the regulatory properties of the enzyme. In addition, the characterization of adenylyl cyclase by mutagenic techniques and the identification of the binding site for P-site inhibitors have led to modeling studies that describe the ATP-binding site. Despite these advances, the catalytic mechanism of adenylyl cyclase remains uncertain, especially with respect to the role that magnesium ions may play in this process. We have identified four mutant mammalian adenylyl cyclases defective in their metal dependence, allowing us to further characterize the function of metal ions in the catalytic mechanism of this enzyme. The wild-type adenylyl cyclase shows a biphasic Mg2+ dose-response curve in which the high-affinity component displays cooperativity (Hill coefficient of 1.4). Two mutations (C441R and Y442H) reduce the affinity of the adenylyl cyclase for Mg2+ dramatically without affecting the binding of MgATP, suggesting that there is a metal requirement in addition to the ATP-bound Mg2+. The results of this study thus demonstrate multiple metal requirements of adenylyl cyclase and support the existence of a Mg2+ ion essential for catalysis and distinct from the ATP-bound ion. We propose that adenylyl cyclase employs a catalytic mechanism analogous to that of DNA polymerase, in which two key magnesium ions facilitate the nucleophilic attack of the 3'-hydroxyl group and the subsequent elimination of pyrophosphate.  相似文献   

20.
DNase I of tilapia (Oreochromis mossambicus) was purified to homogeneity. Tilapia DNase I is most active at pH 8.5 with Mg2+ as activator. The Ca2+/Mg2+ pair has a synergistic effect on activation. The enzyme is readily inactivated by heating above 55 degrees C, but is not inactivated by trypsin or 2-mercaptoethanol under alkaline conditions, with or without CaCl2. Its isoelectric point is 6.0. The 258-amino-acid sequence of tilapia DNase I was derived from overlapping sequences of tryptic, chymotryptic and CNBr peptides. The purified enzyme has two variants differing by a single Lys-->Arg mutation at position 125. The polypeptide chain has one disulfide bridge and one carbohydrate side chain. By mass spectrometry, the purified enzyme shows many molecular mass forms differing by Lys/Arg substitution and sugar-chain length. The major form has a molecular mass of 30,914 Da. A 1061-bp nucleotide sequence for the cDNA of tilapia DNase I, obtained by gene cloning and DNA sequencing, contains an ORF coding for a putative 26-residue transmembrane peptide and the mature DNase I polypeptide.  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-23

京公网安备 11010802026262号