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.
2.
Specific binding sites for vasoactive intestinal peptide were characterized in plasma membranes from rat intestinal epithelial cells. At 30°C, the interaction of 125I-labelled peptide with intestinal membranes was rapid, reversible, specific and saturable. At equilibrium, the binding of 125I-labelled peptide was competitively inhibited by native peptide in the 3 · 10?11?3 · 10?7 M range concentration. Scatchard analysis of binding data suggested the presence of two distinct classes of vasoactive intestinal peptide binding sites: a class with a high affinity Kd = 0.28 nM) and a low capacity (0.8 pmol peptide/mg membrane protein) and a class with a low affinity (Kd = 152 nM) and a high capacity (161 pmol peptide/mg membrane protein). Secretin competitively inhibited binding of 125I-labelled peptide but its potency was 1/1000 that of native peptide. Glucagon and the gastric inhibitory peptide were ineffective. The guanine nucleotides, GTP and Gpp(NH)p inhibited markedly the interaction of 125I-labelled peptide with its binding sites, by increasing the rate of dissociation of peptide bound to membranes. The other nucleotides triphosphate tested (ATP, ITP, UTP, CTP) were also effective in inhibiting binding of 125I-labelled peptide to membranes but their potencies were 1/100-1/1000 that of guanine nucleotides.The specificity and affinity of the vasoactive intestinal peptide-binding sites in plasma membranes prepared from rat intestinal epithelial cells, which is in agreement with an adenylate cyclase highly sensitive to the peptide recently characterized in these membranes (Amiranoff, B., Laburthe, M., Dupont, C. and Rosselin, G. (1978) Biochim. Biophys. Acta 544, 474–481) further argue for a physiological role of the peptide in the regulation of intestinal epithelial function.  相似文献   

3.
We have purified an ATP-dependent protease with protein-dependent ATPase activity from bovine adrenal cortex mitochondria to near homogeneity. The subunit molecular weight is 108,000 and the enzyme appears to be a hexamer with approximately identical subunits. Based on the experiments using various nucleoside triphosphates and their related compounds, it is concluded that hydrolysis of the high-energy bond in nucleoside triphosphates is not an absolute requirement for proteolysis. Nucleotide specificity of this enzyme varies, depending on the protein or peptide substrates used. When casein was the substrate, ATP and dATP were quite effective, but other nucleotides were not. When insulin and angiotensinogen were used as substrate, ATP, other nucleoside triphosphates, ADP, inorganic triphosphate, pyrophosphate, and phosphate were effective. One of the cleaving linkages hydrolyzed by this enzyme was revealed to be the Leu-Leu bond of angiotensinogen. However, the specificity appears to be broad in view of the hydrolysis pattern of glucagon.  相似文献   

4.
The specificity of the ATP:corrinoid adenosyltransferase (CobA) enzyme of Salmonella enterica serovar Typhimurium LT2 for its nucleotide substrate was tested using ATP analogs and alternative nucleotide donors. The enzyme showed broad specificity for the nucleotide base and required the 2'-OH group of the ribosyl moiety of ATP for activity. (31)P NMR spectroscopy was used to identify inorganic triphosphate (PPP(i)) as the byproduct of the reaction catalyzed by the CobA enzyme. Cleavage of triphosphate into pyrophosphate and orthophosphate did not occur, indicating that triphosphate cleavage was not required for release of the adenosylcorrinoid product. Triphosphate was a strong inhibitor of the reaction, with 85% of CobA activity lost when the ATP/PPP(i) ratio present in the reaction mixture was 1:2.5.  相似文献   

5.
Whereas the chemotactic peptide, N-formyl-L-methionyl-L-leucyl-L-phenylalanine (fMet-Leu-Phe), induced NADPH-oxidase-catalyzed superoxide (O2-) formation in human neutrophils, purine and pyrimidine nucleotides per se did not stimulate NADPH oxidase but enhanced O2- formation induced by submaximally and maximally stimulatory concentrations of fMet-Leu-Phe up to fivefold. On the other hand, FMet-Leu-Phe primed neutrophils to generate O2- upon exposure to nucleotides. At a concentration of 100 microM, purine nucleotides enhanced O2- formation in the effectiveness order adenosine 5'-O-[3-thio]triphosphate (ATP[gamma S]) greater than ITP greater than guanosine 5'-O-[3-thio]triphosphate (GTP[gamma S]) greater than ATP = adenosine 5'-O-[2-thio]triphosphate (Sp-diastereomer) = GTP = guanosine 5'-O-[2-thio]diphosphate (GDP[beta S] = ADP greater than adenosine 5'-[beta, gamma-imido]triphosphate = adenosine 5'-O-[2-thio]triphosphate] (Rp-diastereomer). Pyrimidine nucleotides stimulated fMet-Leu-Phe-induced O2- formation in the effectiveness order uridine 5'-O-[3-thio]triphosphate (UTP[gamma S]) = UTP greater than CTP. Uracil (UDP[beta S]) = uridine 5'-O[2-thio]triphosphate (Rp-diastereomer) (Rp)-UTP[beta S]) = UTP greater than CTP. Uracil nucleotides were similarly effective potentiators of O2- formation as the corresponding adenine nucleotides. GDP[beta S] and UDP[beta S] synergistically enhanced the stimulatory effects of ATP[gamma S], GTP[gamma S] and UTP[gamma S]. Purine and pyrimidine nucleotides did not induce degranulation in neutrophils but potentiated fMet-Leu-Phe-induced release of beta-glucuronidase with similar nucleotide specificities as for O2- formation. In contrast, nucleotides per se induced aggregation of neutrophils. Treatment with pertussis toxin prevented aggregation induced by both nucleotides and fMet-Leu-Phe. Our results suggest that purine and pyrimidine nucleotides act via nucleotide receptors, the nucleotide specificity of which is different from nucleotide receptors in other cell types. Neutrophil nucleotide receptors are coupled to guanine-nucleotide-binding proteins. As nucleotides are released from cells under physiological and pathological conditions, they may play roles as intercellular signal molecules in neutrophil activation.  相似文献   

6.
Terminal deoxynucleotidyltransferase is the only DNA polymerase that is strongly inhibited in the presence of ATP. We have labeled calf terminal deoxynucleotidyltransferase with [32P]ATP in order to identify its binding site in terminal deoxynucleotidyltransferase. The specificity of ATP cross-linking to terminal deoxynucleotidyltransferase is shown by the competitive inhibition of the overall cross-linking reaction by deoxynucleoside triphosphates, as well as the ATP analogs Ap4A and Ap5A. Tryptic peptide mapping of [32P]ATP-labeled enzyme revealed a peptide fraction that contained the majority of cross-linked ATP. The properties, chromatographic characteristics, amino acid composition, and sequence analysis of this peptide fraction were identical with those found associated with dTTP cross-linked terminal deoxynucleotidyl-transferase peptide (Pandey, V. N., and Modak, M. J. (1988a). J. Biol. Chem. 263, 3744-3751). The involvement of the same 2 cysteine residues in the crosslinking of both nucleotides further confirmed the unity of the ATP and dTTP binding domain that contains residues 224-237 in the primary amino acid sequence of calf terminal deoxynucleotidyltransferase.  相似文献   

7.
ClpX requires ATP to unfold protein substrates and translocate them into the proteolytic chamber of ClpP for degradation. The steady-state parameters for hydrolysis of ATP and ATPgammaS by ClpX were measured with different protein partners and the kinetics of degradation of ssrA-tagged substrates were determined with both nucleotides. ClpX hydrolyzed ATPgammaS to ADP and thiophosphate at a rate (6/min) significantly slower than ATP hydrolysis (140/min), but the hydrolysis of both nucleotides was increased by ssrA-tagged substrates and decreased by ClpP. K(M) and k(cat) for hydrolysis of ATP and ATPgammaS were linearly correlated over a 200-fold range, suggesting that protein partners largely affect k(cat) rather than nucleotide binding, indicating that most bound ATP leaves the enzyme by hydrolysis rather than dissociation, and placing an upper limit of approximately 15 micro M on K(D) for both nucleotides. Competition studies with ClpX and fluorescently labeled ADP gave inhibition constants for ATPgammaS ( approximately 2 micro M) and ADP ( approximately 3 micro M) under the reaction conditions used for steady-state kinetics. In the absence of Mg(2+), where hydrolysis does not occur, the inhibition constant for ATP ( approximately 55 micro M) was weaker but very similar to the value for ATPgammaS ( approximately 45 micro M). Compared with ATP, ATPgammaS supported slow but roughly comparable rates of ClpXP degradation for two Arc-ssrA substrates and denatured GFP-ssrA, but not of native GFP-ssrA. These results show that the processing of protein substrates by ClpX is closely coupled to the maximum rate of nucleotide hydrolysis.  相似文献   

8.
It was shown that low concentrations of ATP (1..10(-4)M) and 10-fold concentrations of AMP (1.10(-3)M) at three constant L-threonine concentrations activated the L-threonine dehydratase activity of L-threonine-L-serine dehydratase from human liver, but had no effect on the L-serine dehydratase activity of this enzyme. Higher concentrations of both nucleotides inhibited the enzyme. The effects of ATP and AMP were specific. The activating and inhibiting effects of various concentrations of ATP and AMP were revealed as changes in the shapes of the curves for the initial reaction rate of the L-threonine dehydratase reaction versus initial substrate concentration. For this reaction the curves were not hyperbolic and were characterized by intermediary plateaux. ATP and AMP also influenced the maximal rate of the enzymatic reaction. Using the desensitization method it was shown that the activating effects of ATP and AMP are of allosteric nature. Thus, human liver L-threonine-L-serine dehydratase is an allosteric enzyme, for which positive allosteric effectors are low concentrations of ATP and AMP and negative allosteric effectors are high concentrations of these nucleotides. A possible mechanism of allosteric regulation of the enzyme under catalysis of the L-threonine dehydratase reaction and the lack of regulation under catalysis of the L-serine dehydratase reaction as well as specificity of the allosteric sites of this enzyme to the two nucleotides and the physiological significance of this process are discussed.  相似文献   

9.
In order to understand better the structural and functional relations between protein kinase CK2 catalytic subunit, the triphosphate moiety of ATP, the catalytic metal and the peptidic substrate, we built a structural model of Yarrowia lipolytica protein kinase CK2 catalytic subunit using the recently solved three-dimensional structure of the maize enzyme and the structure of cAMP-dependent protein kinase peptidic inhibitor (1CDK) as templates. The overall structure of the catalytic subunit is close to the structure solved by Niefind et al. It comprises two lobes, which move relative to each other. The peptide used as substrate is tightly bound to the enzyme, at specific locations. Molecular dynamic calculations in combination with the study of the structural model led us to identify amino acid residues close to the triphosphate moiety of ATP and a residue sufficiently far from the peptide that could be mutated so as to modify the specificity of the enzyme. Site-directed mutagenesis was used to replace by charged residues both glycine-48, a residue located within the glycine-rich loop, involved in binding of ATP phosphate moiety, and glycine-177, a residue close to the active site. Kinetic properties of purified wild-type and mutated subunits were studied with respect to ATP, MgCl(2) and protein kinase CK2 specific peptide substrates. The catalytic efficiency of the G48D mutant increased by factors of 4 for ATP and 17.5 for the RRRADDSDDDDD peptide. The mutant G48K had a low activity with ATP and no detectable activity with peptide substrates and was also inhibited by magnesium. An increased velocity of ADP release by G48D and the building of an electrostatic barrier between ATP and the peptidic substrate in G48K could explain these results. The kinetic properties of the mutant G177K with ATP were not affected, but the catalytic efficiency for the RRRADDSDDDDD substrate increased sixfold. Lysine 177 could interact with the lysine-rich cluster involved in the specificity of protein kinase CK2 towards acidic substrate, thereby increasing its activity.  相似文献   

10.
A J Zaug  C A Grosshans  T R Cech 《Biochemistry》1988,27(25):8924-8931
A shortened form of the self-splicing intervening sequence RNA of Tetrahymena acts as a sequence-specific endoribonuclease. Specificity of cleavage is determined by Watson-Crick base pairing between the active site of the RNA enzyme (ribozyme) and its RNA substrate [Zaug, A. J., Been, M. D., & Cech, T. R. (1986) Nature (London) 324, 429-433]. Surprisingly, single-base changes in the substrate RNA 3 nucleotides preceding the cleavage site, giving a mismatched substrate-ribozyme complex, enhance the rate of cleavage. Mismatched substrates show up to a 100-fold increase in kcat and, in some cases, in kcat/Km. A mismatch introduced by changing a nucleotide in the active site of the ribozyme has a similar effect. Addition of 2.5 M urea or 3.8 M formamide or decreasing the divalent metal ion concentration from 10 to 2 mM reverses the substrate specificity, allowing the ribozyme to discriminate against the mismatched substrate. The effect of urea is to decrease kcat and kcat/Km for cleavage of the mismatched substrate; Km is not significantly affected at 0-2.5 M urea. Thus, progressive destabilization of ribozyme-substrate pairing by mismatches or by addition of a denaturant such as urea first increases the rate of cleavage to an optimum value and then decreases the rate.  相似文献   

11.
The active site glutamate (Glu(111)) and the active site histidine (His(112)) of insulin-degrading enzyme (IDE) were mutated. These mutant enzymes exhibit, in addition to a large decrease in catalytic activity, a change in the substrate-velocity response from a sigmoidal one seen with the native enzyme (Hill coefficient > 2), to a hyperbolic response. With 2-aminobenzoyl-GGFLRKHGQ-N-(2,4-dinitrophenyl)ethylenediamine as substrate, ATP and triphosphate increase the reaction rate of the wild type enzyme some 50-80-fold. This effect is dampened with glutamate mutants to no effect or less than a 3-fold increase in activity and changed to inhibition with the histidine mutants. Sedimentation equilibrium shows the IDE mutants exhibit a similar oligomeric distribution as the wild type enzyme, being predominantly monomeric, with triphosphate having little if any effect on the oligomeric state. Triphosphate did induce aggregation of many of the IDE mutants. Thus, the oligomeric state of IDE does not correlate with kinetic properties. The His(112) mutants were shown to bind zinc, but with a lower affinity than the wild type enzyme. The glutamate mutants displayed an altered cleavage profile for the peptide beta-endorphin. Wild type IDE cleaved beta-endorphin at Leu(17)-Phe(18) and Phe(18)-Lys(19), whereas the glutamate mutants cleaved at these sites, but in addition at Lys(19)-Asn(20) and at Met(5)-Thr(6). Thus, active site mutations of IDE are suggested to not only reduce catalytic activity but also cause local conformational changes that affect the allosteric properties of the enzyme.  相似文献   

12.
Protease La is an ATP-dependent protease that catalyzes the rapid degradation of abnormal proteins and certain normal polypeptides in Escherichia coli. In order to learn more about its specificity and the role of ATP, we tested whether small fluorogenic peptides might serve as substrates. In the presence of ATP and Mg2+, protease La hydrolyzes two oligopeptides that are also substrates for chymotrypsin, glutaryl-Ala-Ala-Phe-methoxynaphthylamine (MNA) and succinyl-Phe-Leu-Phe-MNA. Methylation or removal of the acidic blocking group prevented hydrolysis. Closely related peptides (glutaryl-Gly-Gly-Phe-MNA and glutaryl-Ala-Ala-Ala-MNA) are cleaved only slightly, and substrates of trypsin-like proteases are not hydrolyzed. Furthermore, several peptide chloromethyl ketone derivatives that inhibit chymotrypsin and cathepsin G (especially benzyloxycarbonyl-Gly-Leu-Phe-chloro-methyl ketone), inhibited protease La. Thus its active site prefers peptides containing large hydrophobic residues, and amino acids beyond the cleavage site influence rates of hydrolysis. Peptide hydrolysis resembles protein breakdown by protease La in many respects: 1) ADP inhibits this process rapidly, 2) DNA stimulates it, 3) heparin, diisopropyl fluorophosphate, and benzoyl-Arg-Gly-Phe-Phe-Leu-MNA inhibit hydrolysis, 4) the reaction is maximal at pH 9.0-9.5, 5) the protein purified from lon- E. coli or Salmonella typhymurium showed no activity against the peptide, and that from lonR9 inhibited peptide hydrolysis by the wild-type enzyme. With partially purified enzyme, peptide hydrolysis was completely dependent on ATP. The pure protease hydrolyzed the peptide slowly when only Mg2+, Ca2+, or Mn2+ were present, and ATP enhanced this activity 6-15-fold (Km = 3 microM). Since these peptides cannot undergo phosphorylation, adenylylation, modification of amino groups, or denaturation, these mechanisms cannot account for the stimulation by ATP. Most likely, ATP and Mg2+ affect the conformation of the enzyme, rather than that of the substrate.  相似文献   

13.
Phosphoenolpyruvate carboxykinases, depending on the enzyme origin, preferentially use adenine or guanine nucleotides as substrates. In this work, analyses of the substrate specificity of the Saccharomyces cerevisiae ATP-dependent enzyme have been carried out. Kinetics studies gave relative values of k(cat)/K(m) for the nucleoside triphosphate complexes in the order ATP>GTP>ITP>UTP>CTP. For the nucleoside diphosphate complexes the order is ADP>GDP>IDP congruent withUDP>CDP. This shows that the enzyme has a strong preference for ADP (or ATP) over other nucleotides, being this preference about an order of magnitude higher for the diphosphorylated than for the triphosphorylated nucleosides. The calculated binding free energies (kcalmol(-1)) at 25 degrees C are 7.39 and 6.51 for ATP and ADP, respectively. These values decrease with the nucleotide structure in the same order than the kinetic specificity. The binding energy for any triphosphorylated nucleoside is more favourable than for the corresponding diphosphorylated compound, showing the relevance of the P(gamma) for nucleotide binding. Homology models of the adenine and guanine nucleotides in complex with the enzyme show that the base adopts a similar conformation in the diphosphorylated nucleosides while in the triphosphorylated nucleosides the sugar-base torsion angle is 61 degrees for ATP and -53 degrees for GTP. Differences are also noted in the distance between P(beta) and Mn2+ at site 1. This distance is almost the same in the ATP, GTP, and UTP complexes, however in the ADP, GDP and UDP complexes it is 2.9, 5.1, and 7A, respectively. Experimental data obtained with a Thr463Ala mutant enzyme agree with molecular simulation predictions. The results here presented are discussed in terms of the proposed interactions of the nucleotides with the protein.  相似文献   

14.
The mechanism of regulation of actomyosin subfragment 1 ATPase   总被引:9,自引:0,他引:9  
The mechanism of regulation of actin-subfragment 1 nucleoside triphosphatase is described in terms of the rate and equilibrium constants of a relatively simple kinetic scheme: (Formula: see text) where T, D, and Pi are nucleoside triphosphate, nucleoside diphosphate, and inorganic phosphate, respectively; Ka, Kb, and Kc are association constants; the ki are first-order rate constants; A is regulated actin (actin-tropomyosin-troponin); and M is subfragment 1. Calcium binding to regulated actin had little effect on step 2; k2 was almost unaffected, and k-2 increased, at most, 2-fold. k-1 and k3 increased 10-20-fold for ATP and 3-5-fold for 1-N6-ethenoadenosine triphosphate as substrates. Kb and Kc increased by less than 50%, whereas Ka increased 6-10-fold. The primary effect in regulation is on the rate of a conformational change which determines the rate of dissociation of ligands bound to the active site. The measurements probably underestimate the ratio of rate constants of product dissociation for active and relaxed states of actin because of heterogeneity. The kinetic evidence can be explained by a partial steric blocking mechanism or by a conformational (nonsteric) mechanism.  相似文献   

15.
To localize and characterize the regulatory nucleotide site of skeletal muscle sarcoplasmic reticulum Ca2+-ATPase, we have investigated the effects of ADP, ATP, and analogues of these nucleotides on the rate of dephosphorylation of both native ATPase and ATPase modified with fluorescein 5'-isothiocyanate (FITC), a reagent which hinders access of nucleotides to the ATPase catalytic site without affecting phosphorylation from Pi. Dephosphorylation of the phosphoenzyme formed from Pi was monitored by rapid filtration or stopped-flow fluorescence, mostly at 20 degrees C, pH 6.0, and in the absence of potassium. Fluorescence measurements were made possible through the use of 8-bromo-ATP, which selectively quenched certain tryptophan residues of the ATPase, thereby allowing the intrinsic fluorescence changes associated with dephosphorylation to be measured in the presence of bound nucleotide. ATP, 8-bromo-ATP, and trinitrophenyladenosine diand triphosphate, but not ADP, enhanced the rate of dephosphorylation of native ATPase 2-3-fold when added in the absence of divalent cations. Millimolar concentrations of Mg2+ eliminated the accelerating effects. Acceleration in the absence of Mg2+ was observed at relatively low concentrations of ATP and 8-bromo-ATP (0.01-0.1 mM) and binding of metal-free ATP and ADP, but not Mg.ATP, to the phosphoenzyme in this concentration range was demonstrated directly. Modification of the ATPase with FITC blocked nucleotide binding in the submillimolar concentration range and eliminated the nucleotide-induced acceleration of dephosphorylation. These results show that dephosphorylation, under these conditions, is regulated by ATP but not by Mg.ATP or ADP, and that the catalytic site is the locus of this "regulatory" ATP binding site.  相似文献   

16.
The expression of certain COOH-terminal truncation mutants of the epidermal growth factor receptor (EGFR) can lead to cell transformation, and with ligand stimulation, a broader spectrum of phosphorylated proteins appears compared with EGF-treated cells expressing wild-type EGFR. Accordingly, it has been proposed that elements within the COOH terminus may determine substrate specificity of the EGFR tyrosine kinase (Decker, S. J., Alexander, C., and Habib, T. (1992) J. Biol. Chem. 267, 1104-1108; Walton, G. M., Chen, W. S., Rosenfeld, M. G., and Gill, G. N. (1990) J. Biol. Chem. 265, 1750-1754). To address this hypothesis, we analyzed in vitro the steady-state kinetic parameters for phosphorylation of several substrates by both wild-type EGFR and an oncogenic EGFR mutant (the ct1022 mutant) truncated at residue 1022. The substrates included: (i) a phospholipase C-gamma fragment (residues 530-850); (ii) the 46-kDa isoform of the Shc adapter protein; (iii) a 13-residue peptide mimic for the region around the major autophosphorylation tyrosine and the Shc binding site (the Y1173 peptide); (iv) a poly(Glu,Tyr) 4:1 copolymer; and (v) the 8-residue peptide, angiotensin II. Our data demonstrate that the steady-state kinetic parameters for the ct1022 mutant differ from those of the wild-type enzyme, and the differences are substrate-dependent. These results support the concept that this oncogenic truncation/mutation alters EGFR substrate specificity, rather than causing a general alteration of activity. We performed the experiments using a non-radioactive fluorescence polarization assay that quantifies the degree of phosphorylation of peptide as well as natural substrates. The results are consistent with those from the traditional [gamma-32P]ATP/filtration assay.  相似文献   

17.
R Mahmood  M Elzinga  R G Yount 《Biochemistry》1989,28(9):3989-3995
A portion of the active site of rabbit skeletal myosin near the ribose ring of ATP can be labeled by the photoaffinity analogue 3'(2')-O-(4-benzoylbenzoyl)adenosine triphosphate (Bz2ATP). The specificity of the photolabeling was assured by first trapping [14C]Bz2ATP at the active site by use of thiol cross-linking agents [Mahmood, R., Cremo, C., Nakamaye, K., & Yount, R. (1987) J. Biol. Chem. 262, 14479-14486]. Five radioactive peptides were isolated by high-performance liquid chromatography after extensive trypsin and subtilisin digestion of photolabeled myosin subfragment 1. Four of these peptides were sequenced by Edman techniques, and all originated from a region with the sequence Gly-Glu-Ile-Thr-Val-Pro-Ser-Ile-Asp-Asp-Gln, which corresponds to rabbit myosin heavy chain residues 318-328. The fifth labeled peptide had an amino acid composition appropriate for residues 312-328. Amino acid composition, radiochemical analysis, and sequence data indicate that Ser-324 is the major amino acid residue photolabeled by Bz2ATP. Spectrophotometric evidence indicates that the benzophenone carbonyl group has inserted into a C-H bond from either the alpha- or beta-carbon of serine. These results place Ser-324 at a distance of 6-7 A from the 3'(2') ribose oxygens of ATP bound at the active site of myosin.  相似文献   

18.
The interactions of nucleotides at the allosteric and catalytic sites of phosphorylase kinase were examined. Binding of nucleoside triphosphates at the nucleoside diphosphate allosteric activation site inhibited enzymatic activity; this was observed with either ATP or GTP. Increasing concentrations of ADP caused a biphasic response: low concentrations activated and higher concentrations inhibited. Inhibition was due to the binding of ADP at the catalytic site, as opposed to an allosteric inhibitory site. GDP activated at low concentrations, but did not inhibit even at relatively high concentrations, and is therefore a specific probe for the allosteric site. Maximal activity of the nonactivated holoenzyme at pH 6.8 is achieved at an optimal ratio of ATP to ADP, such that the inhibitory actions of ATP at the allosteric site and of ADP at the catalytic site are balanced. Various potential molecular mechanisms to explain the allosteric activation by ADP were examined and ruled out, thus strengthening our previous conclusion that the activation is predominantly caused by a conformational transition in the beta subunits directly induced by the binding of ADP (Cheng, A., Fitzgerald, T. J., and Carlson, G. M. (1985) J. Biol. Chem. 260, 2535-2542; Trempe, M. R., and Carlson, G. M. (1987) J. Biol. Chem. 262, 4333-4340; Cheng, A., Fitzgerald, T. J., Bhatnager, D., Roskoski, R., Jr., and Carlson, G. M. (1988) J. Biol. Chem. 263, 5534-5542). The catalytic site exhibited high stereospecificity for inhibition by the Rp and Sp epimers of adenosine 5'-O-(1-thiodiphosphate), with the Rp epimer (Ki = 0.5 microM) being 136-fold more effective than its Sp counterpart. This can readily explain the inability of the Rp epimer to be an effective allosteric activator.  相似文献   

19.
Escherichia coli RNase T, the enzyme responsible for the end-turnover of tRNA and for the 3' maturation of 5 S and 23 S rRNAs and many other small, stable RNAs, was examined in detail with respect to its substrate specificity. The enzyme was found to be a single-strand-specific exoribonuclease that acts in the 3' to 5' direction in a non-processive manner. However, although other Escherichia coli exoribonucleases stop several nucleotides downstream of an RNA duplex, RNase T can digest RNA up to the first base pair. The presence of a free 3'-hydroxyl group is required for the enzyme to initiate digestion. Studies with RNA homopolymers and a variety of oligoribonucleotides revealed that RNase T displays an unusual base specificity, discriminating against pyrimidine and, particularly, C residues. Although RNase T appears to bind up to 10 nucleotides in its active site, its specificity is defined largely by the last 4 residues. A single 3'-terminal C residue can reduce RNase T action by >100-fold, and 2-terminal C residues essentially stop the enzyme. In vivo, the substrates of RNase T are similar in that they all contain a double-stranded stem followed by a single-stranded 3' overhang; yet, the action of RNase T on these substrates differs. The substrate specificity described here helps to explain why the different substrates yield different products, and why certain RNA molecules are not substrates at all.  相似文献   

20.
A synthetic peptide analog of the precursor region of preproparathyroid hormone has been shown to be a specific substrate for hen oviduct signal peptidase. The sequence of the 31-residue peptide is Ser-Ala-Lys-Asp-norleucine (Nle)-Val-Lys-Val-Nle-Ile-Val-Nle-Leu-Ala-Ile-Ala-Phe-Leu-Ala-Arg-Ser-As p-Gly-Lys-Ser-Val-Lys-Lys-Arg-D-Tyr-amide (Caulfield, M. P., Duong, L. T., O'Brien, R., Majzoub, J. A., and Rosenblatt, M. (1988) Mol. Endocrinol. 2, 452-458). This sulfur-free signal peptide analog can be labeled with 125I on the C-terminal D-tyrosine and is cleaved by purified hen oviduct signal peptidase between Gly and Lys, the correct site of cleavage of preproparathyroid hormone in vivo. Amino acid sequence analysis of the cleavage product released 125I at the seventh cycle of Edman degradation, confirming that enzymatic cleavage occurs at the physiological site. Synthetic peptide analogs of the substrate with Lys, Pro, or Asp substituted for Nle-18 were poor substrates for the enzyme and were also poor competitive inhibitors of catalysis, suggesting that modifications at position -18, 12 amino acids from the site of cleavage, directly influence binding by the enzyme. Analysis of the reactivity of signal peptidase with these synthetic peptides provides insight into the cleavage specificity requirements of this eukaryotic signal peptidase.  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