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.
Four tight nucleotide binding sites of chloroplast coupling factor 1.   总被引:1,自引:0,他引:1  
We have examined the properties of the four tight nucleotide binding sites of reductively activated chloroplast coupling factor 1. Tight sites are here defined as those which retain bound nucleotides after passage of the chloroplast coupling factor 1 through Sephadex gel filtration centrifuge columns. Two of the sites, here called sites 4 and 5, have not been characterized in detail before. Site 4 has properties similar to those of site 1. It binds to ADP, ATP, and adenylyl-beta,gamma-imidodiphosphate (AMP-PNP) tightly in the presence or absence of Mg2+. Bound ADP exchanges rapidly with medium ADP, but rapid exchange with ATP or AMP-PNP requires Mg2+. Site 4 may slowly hydrolyze bound ATP in the absence of medium nucleotides. Site 5 has properties similar to those of site 2. Tight binding of ATP and AMP-PNP requires Mg2+, but Mg29+)-ADP is not tightly bound. Site 5 does not hydrolyze bound ATP in the absence of medium nucleotides. Complete filling of all four tight nucleotide binding sites requires about one millimolar nucleotide, suggesting that low affinity binding sites are converted to tight binding via a nucleotide binding-induced conformational change.  相似文献   

2.
The mechanism of assisted protein folding by the chaperonin GroEL alone or in complex with the co-chaperonin GroES and in the presence or absence of nucleotides has been subject to extensive investigations during the last years. In this paper we present data where we have inactivated GroEL by stepwise blocking the nucleotide binding sites using the non-hydrolyzable ATP analogue, (Cr(H2O)4)3+ATP. We correlated the amount of accessible nucleotide binding sites with the residual ATP hydrolysis activity of GroEL as well as the residual refolding activity for two different model substrates. Under the conditions used, folding of the substrate proteins and ATP hydrolysis were directly proportional to the residual, accessible nucleotide binding sites. In the presence of GroES, 50% of the nucleotide binding sites were protected from inactivation by CrATP and the resulting protein retains 50% of both ATPase and refolding activity. The results strongly suggest that under the conditions used in our experiments, the nucleotide binding sites are additive in character and that by blocking of a certain number of binding sites a proportional amount of ATP hydrolysis and refolding activities are inactivated. The experiments including GroES suggest that full catalytic activity of GroEL requires both rings of the chaperonin. Blocking of the nucleotide binding sites of one ring still allows function of the second ring.  相似文献   

3.
Characterization of recombinant murine leukemia virus integrase.   总被引:6,自引:6,他引:0       下载免费PDF全文
Retroviral integration involves two DNA substrates that play different roles. The viral DNA substrate is recognized by virtue of specific nucleotide sequences near the end of a double-stranded DNA molecule. The target DNA substrate is recognized at internal sites with little sequence preference; nucleosomal DNA appears to be preferred for this role. Despite this apparent asymmetry in the sequence, structure, and roles of the DNA substrates in the integration reaction, the existence of distinct binding sites for viral and target DNA substrates has been controversial. In this report, we describe the expression in Escherichia coli and purification of Moloney murine leukemia virus integrase as a fusion protein with glutathione S-transferase, characterization of its activity by using several model DNA substrates, and the initial kinetic characterization of its interactions with a model viral DNA substrate. We provide evidence for functionally and kinetically distinct binding sites for viral and target DNA substrates and describe a cross-linking assay for DNA binding at a site whose specificity is consistent with the target DNA binding site.  相似文献   

4.
Two adenine nucleotide binding sites of chloroplast coupling factor 1 (CF1) were shown previously to switch their properties after exposure of the enzyme to Mg2(+)-ATP or Ca2(+)-ATP (Shapiro, A. B., and McCarty, R. E. (1988) J. Biol. Chem. 263, 14160-14165). The change in binding properties was monitored by fluorescence resonance energy transfer between Lucifer Yellow vinyl sulfone covalently bound to one alpha subunit and trinitrophenyl-ATP (TNP-ATP) tightly bound to nucleotide binding site 1. When the nucleotide binding properties of sites 1 and 3 switch during catalysis, site 3, which is nearer Lucifer Yellow than site 1, switches its nucleotide binding properties with site 1, allowing TNP-ATP to become tightly bound near Lucifer Yellow. The smaller separation allows energy transfer to occur, resulting in decreased Lucifer Yellow fluorescence. In this paper, we show that adenylyl-beta,gamma-imidodiphosphate (AMP-PNP) bound to CF1 and caused nucleotide binding sites 1 and 3 to switch properties, but was not hydrolyzed. Using AMP-PNP, we also found that relaxation of the properties of the sites to the precatalysis state after removal of substrate occurred in the absence of hydrolysis of the last bound nucleotide. When Mg2+ was omitted during exposure of CF1 to ATP, there was very little hydrolysis or nucleotide site switching. When Mg2+ was added to a very low concentration which was more than stoichiometric with CF1, however, site switching occurred at its maximal level with virtually no increase in ATP hydrolysis. These results support a model in which binding of substrate Mg2(+)-ATP, not hydrolysis, causes the putative catalytic sites to switch properties, in agreement with the alternating site catalytic cooperativity hypothesis (Boyer, P. D. (1989) FASEB J. 3, 2164-2178). TNP-ATP, the fluorescence acceptor, did not cause nucleotide site switching when incubated with CF1 in the presence of EDTA to eliminate free Mg2+. Two possible additional nucleotide binding sites were detected, in addition to the three well characterized sites. At least one of these sites was close to the Lucifer Yellow site, judging by the amount of energy transfer caused by partial occupancy with TNP-ATP.  相似文献   

5.
Three conformations of the thermosome, an archaeal group II chaperonin, have been determined by cryo-electron microscopy (EM). We describe an open form of the double-ring oligomer, a closed form and a bullet-shaped form with one ring open and the other closed. Domain movements have been deduced by docking atomic coordinates into the EM maps. The subunit apical domains, bearing the putative substrate binding sites, rotate about 30 degrees upwards and twist in the plane of the ring from the closed to the open conformation. The closed rings have their nucleotide binding pockets closed by the intermediate domains, but in the open rings, the pocket is accessible.  相似文献   

6.
The reaction of [Pt(dien)Cl1Cl (dien = NH2CH2CH2NHCH2CH2NH2) with nucleotides has been studied by nuclear magnetic resonance. It has been found that the CMP (cytidine 5'-monophosp-ate) and GMP (guanosine 5'-monophosphate/coordinate to the platinum atom through N3 and N7, respectively. The reaction of the platinum salt with the nucleotide is complete when one to one ratio of platinum to nucleotide is used and no evidence of phosphate group binding to platinum has been found. No additional binding sites have been detected except the N7 site on the guanylic group of GMP even in the presence of a large excess of [Pt(dien) Cl1Cl. The AMP (adenosine 5'monophosphate] coordinates to the platinum at the N1 and/or N7 sites. The reaction of AMP and platinum is complete is complete at a ratio of four platinum to one AMP.  相似文献   

7.
Binding of ADP to rat brain hexokinase provided protection against inactivation of the enzyme by glutaraldehyde or by chymotryptic digestion. Graphical analysis of the inactivation experiments was, in both cases, consistent with the existence of a single ADP binding site and a Kd ≈ 3mM for the hexokinase-ADP complex. Both Cibacron Blue F3GA and tetraiodofluorescein, previously found to have a general affinity for nucleotide binding sites, were competitive (vs. ATP) inhibitors of the enzyme, suggesting that they bound only to the site occupied by the nucleotide substrate, ATP. While alternate interpretations cannot be excluded, it is felt that these results are most consistent with the view that there is a single nucleotide binding site on the enzyme. They thereby may serve to stimulate a search for alternative explanations for the complex inhibitory pattern of ADP which had previously been attributed to the existence of two ADP binding sites on the enzyme (J. Ning, D.L. Purich, and H.J. Fromm, J. Biol. Chem. 244, 3840–3846 (1969).  相似文献   

8.
The amidotransferase family of enzymes utilizes the ammonia derived from the hydrolysis of glutamine for a subsequent chemical reaction catalyzed by the same enzyme. The ammonia intermediate does not dissociate into solution during the chemical transformations. A well-characterized example of the structure and mechanism displayed by this class of enzymes is provided by carbamoyl phosphate synthetase (CPS). Carbamoyl phosphate synthetase is isolated from Escherichia coli as a heterodimeric protein. The smaller of the two subunits catalyzes the hydrolysis of glutamine to glutamate and ammonia. The larger subunit catalyzes the formation of carbamoyl phosphate using 2 mol of ATP, bicarbonate, and ammonia. Kinetic investigations have led to a proposed chemical mechanism for this enzyme that requires carboxy phosphate, ammonia, and carbamate as kinetically competent reaction intermediates. The three-dimensional X-ray crystal structure of CPS has localized the positions of three active sites. The nucleotide binding site within the N-terminal half of the large subunit is required for the phosphorylation of bicarbonate and subsequent formation of carbamate. The nucleotide binding site within the C-terminal domain of the large subunit catalyzes the phosphorylation of carbamate to the final product, carbamoyl phosphate. The three active sites within the heterodimeric protein are separated from one another by about 45 A. The ammonia produced within the active site of the small subunit is the substrate for reaction with the carboxy phosphate intermediate that is formed in the active site found within the N-terminal half of the large subunit of CPS. Since the ammonia does not dissociate from the protein prior to its reaction with carboxy phosphate, this intermediate must therefore diffuse through a molecular tunnel that connects these two sites with one another. Similarly, the carbamate intermediate, initially formed at the active site within the N-terminal half of the large subunit, is the substrate for phosphorylation by the ATP bound to the active site located in the C-terminal half of the large subunit. A molecular passageway has been identified by crystallographic methods that apparently facilitates diffusion between these two active sites within the large subunit of CPS. Synchronization of the chemical transformations is controlled by structural perturbations among the three active sites. Molecular tunnels between distant active sites have also been identified in tryptophan synthase and glutamine phosphoribosyl pyrophosphate amidotransferase and are likely architectural features in an expanding list of enzymes.  相似文献   

9.
Lead cleavage sites in the core structure of group I intron-RNA.   总被引:5,自引:4,他引:1       下载免费PDF全文
Self-splicing of group I introns requires divalent metal ions to promote catalysis as well as for the correct folding of the RNA. Lead cleavage has been used to probe the intron RNA for divalent metal ion binding sites. In the conserved core of the intron, only two sites of Pb2+ cleavage have been detected, which are located close to the substrate binding sites in the junction J8/7 and at the bulged nucleotide in the P7 stem. Both lead cleavages can be inhibited by high concentrations of Mg2+ and Mn2+ ions, suggesting that they displace Pb2+ ions from the binding sites. The RNA is protected from lead cleavage by 2'-deoxyGTP, a competitive inhibitor of splicing. The two major lead induced cleavages are both located in the conserved core of the intron and at phosphates, which had independently been demonstrated to interact with magnesium ions and to be essential for splicing. Thus, we suggest that the conditions required for lead cleavage occur mainly at those sites, where divalent ions bind that are functionally involved in catalysis. We propose lead cleavage analysis of functional RNA to be a useful tool for mapping functional magnesium ion binding sites.  相似文献   

10.
J Heitman  P Model 《The EMBO journal》1990,9(10):3369-3378
The EcoRI restriction endonuclease cleaves DNA molecules at the sequence GAATTC. We devised a genetic screen to isolate EcoRI mutants with altered or broadened substrate specificity. In vitro, the purified mutant enzymes cleave both the wild-type substrate and sites which differ from this by one nucleotide (EcoRI star sites). These mutations identify four residues involved in substrate recognition and catalysis that are different from the amino acids proposed to recognize the substrate based on the EcoRI-DNA co-crystal structure. In fact, these mutations suppress EcoRI mutants altered at some of the proposed substrate binding residues (R145, R200). We argue that these mutations permit cleavage of additional DNA sequences either by perturbing or removing direct DNA-protein interactions or by facilitating conformational changes that allosterically couple substrate binding to DNA scission.  相似文献   

11.
High affinity binding sites for serine and aspartate have been characterized in membranes from Salmonella typhimurium and Escherichia coli. Greater than 80% of these sites have been identified as chemotaxis receptors. Mutants lacking binding sites for these amino acids have been shown to have corresponding defects in taxis. The substrate specificity of each of the receptors in Salmonella is very high; most analogs of serine and aspartate do not bind to these receptor sites and do not affect chemotaxis. The transport of these amino acids is apparently not related to chemotaxis. At least 2500 serine receptors and 1200 aspartate receptors with dissociation constants of about 5 microM are present in the membrane fraction of logarithmically growing cells.  相似文献   

12.
The activation of bovine thyroid adenylate cyclase (ATP pyrophosphate-lyase (cyclizing), EC 4.6.1.1) by Gpp(NH)p has been studied using steady-state kinetic methods. This activation is complex and may be characterized by two Gpp(NH)p binding sites of different affinities with measured constants: Ka1 = 0.1 micro M and Ka2 = 2.9 micro M. GDP beta S does not completely inhibit the Gpp(NH)p activation: analysis of the data is consistent with a single GDP beta S inhibitory site which is competitive with the weaker Gpp(NH)p site. Guanine nucleotide effects upon F- activation of adenylate cyclase have been studied. When App(NH)p is the substrate, 10 micro M GTP along with 10 mM NaF gives higher activity than NaF alone, while GDP together with NaF inhibits the activity by 50% relative to NaF. These features are not observed when the complex is assayed with ATP in the presence of a nucleotide regenerating system or when analogs Gpp)NH)p or GDP beta S are used along with NaF. These effects were studied in three other membrane systems using App(NH)p as substrate: rat liver, rat ovary and turkey erythrocyte. No consistent pattern of guanine nucleotide effects upon fluoride activation could be observed in the different membrane preparations. Previous experiments showed that the size of soluble thyroid adenylate cyclase changed whether membranes were preincubated with Gpp(NH)p or NaF. This size change roughly corresponded to the molecular weight of the nucleotide regulatory protein. This finding, coupled with the present data, suggests that two guanine nucleotide binding sites may be involved in regulating thyroid cyclase and that these sites may be on different protein chains.  相似文献   

13.
Escherichia coli ClpA is a AAA+ (ATPase Associated with diverse cellular Activities) chaperone that catalyzes the ATP‐dependent unfolding and translocation of substrate proteins targeted for degradation by a protease, ClpP. ClpA hexamers associate with one or both ends of ClpP tetradecamers to form ClpAP complexes. Each ClpA protomer contains two nucleotide‐binding sites, NBD1 and NBD2, and self‐assembly into hexamers is thermodynamically linked to nucleotide binding. Despite a number of studies aimed at characterizing ClpA and ClpAP‐catalyzed substrate unfolding and degradation, respectively, to date the field is unable to quantify the concentration of ClpA hexamers available to interact with ClpP for any given nucleotide and total ClpA concentration. In this work, sedimentation velocity studies are used to quantitatively examine the self‐assembly of a ClpA Walker B variant in the presence of ATP. In addition to the hexamerization, we observe the formation of a previously unreported ClpA dodecamer in the presence of ATP. Further, we report apparent equilibrium constants for the formation of each ClpA oligomer obtained from direct boundary modeling of the sedimentation velocity data. The energetics of nucleotide binding to NBD1 and NBD2 are revealed by examining the dependence of the apparent association equilibrium constants on free nucleotide concentration.  相似文献   

14.
recA protein promotes homologous pairing and strand exchange by an ordered reaction in which the protein first polymerizes on single-stranded DNA. This presynaptic intermediate, which can be formed either in the presence or absence of Escherichia coli single-stranded binding protein (SSB), has been isolated by gel filtration and characterized. At saturation, purified complexes contained one molecule of recA protein per 3.6 nucleotide residues of single-stranded DNA. Complexes that had been formed in the presence of SSB contained up to one molecule of SSB per 15 nucleotide residues, but the content of SSB in different preparations of isolated complexes appeared to be inversely related to the content of recA protein. Even when they have lost as much as a third of their recA protein, presynaptic complexes can retain activity, because the formation of stable joint molecules depends principally on the binding of recA protein to the single-stranded DNA in the localized region that corresponds to the end of the duplex substrate.  相似文献   

15.
Kidney alkaline phosphatase is an enzyme which requires two types of metals for maximal activity: zinc, which is essential, and magnesium, which is stimulatory. The main features of the Mg2+ stimulation have been analyzed. The stimulation is pH-dependent and is observed mainly between pH 7.5 and 10.5. Mg2+ binding to native alkaline phosphatase is characterized by a dissociation constant of 50 muM at pH 8.5,25 degrees. Binding of Zn2+ is an athermic process. Both the rate constants of association, ka, and of dissociation, kd, have low values. Typical values are 7 M(-1) at pH 8.0, 25 degrees, for ka and 4.10(-4) S(-1) at pH 8.0, 25 degrees, for kd. The on and off processes have high activation energies of 29 kcal mol (-1). Mg2+ can be replaced at its specific site by Mn2+, Co2+, Ni2+, and Zn2+. Zinc binding to the Mg2+ site inhibits the native alkaline phosphatase. Mn2+, Co2+, and Ni2+ also bind to the Mg2+ site with a stimulatory effect which is nearly identic-al with that of Mg2+, Mn2+ is the stimulatory cation which binds most tightly to the Mg2+ site; the dissociation constant of the Mn2+ kidney phosphatase complex is 2 muM at pH 8.5. The stoichiometry of Mn2+ binding has been found to be 1 eq of Mn2+ per mol of dimeric kidney phosphatase. The native enzyme displays absolute half-site reactivity for Mn2+ binding. Mg2+ binding site and the substrate binding sites are distinct sites. The Mg2+ stimulation corresponds to an allosteric effect. Mg2+ binding to its specific sites does not affect substrate recognition, it selectively affects Vmax values. Quenching of the phosphoenzyme formed under steady state conditions with [32P]AMP as a substrate as well as stopped flow analysis of the catalyzed hydrolysis of 2,4-dinitrophenyl phosphate or p-nitrophenyl phosphate have shown that the two active sites of the native and of the Mg2+-stimulated enzyme are not equivalent. Stopped flow analysis indicated that one of the two active sites was phosphorylated very rapidly whereas the other one was phosphorylated much more slowly at pH 4.2. Half of the sites were shown to be reactive at pH 8.0. Quenching experiments have shown that only one of the two sites is phosphorylated at any instant; this result was confirmed by the stopped flow observation of a burst of only 1 mol of nitrophenol per mol of dimeric phosphatase in the pre-steady state hydrolysis of p-nitrophenyl phosphate. The half-of-the-sites reactivity observed for the native and for the Mg2+-stimulated enzyme indicates that the same type of complex, the monophosphorylated complex, accumulates under steady state conditions with both types of enzymes. Mg2+ binding to the native enzyme at pH 8.0 increases considerably the dephosphorylation rate of this monophosphorylated intermediate. A possible mechanism of Mg2+ stimulation is discussed.  相似文献   

16.
P Davanloo  D M Crothers 《Biochemistry》1976,15(20):4433-4438
A method is reported for measuring the stoichiometry of complex formation between actinomycin and a series of deoxynucleotides. The amount of bound actinomycin is measured by distribution of the drug between two liquid phases, a buffer phase containing deoxynucleotide and an organic phase in which the nucleotide is insoluble. Using simple statistical mechanical analysis, the equilibrium equations for several models of actinomycin-deoxynucleotide complexes have been derived: actinomycin with one binding site, with two equivalent independent binding sites, and with two sites which must be occupied together. The binding of actinomycin C3 with dpG, dpApG, dpA, and dpGpC has been examined compared with these models. It is found that binding to dpG and dpApG involves two independent binding sites of nearly equal affinity for nucleotides, whereas binding of dpGpC to the two binding sites on actinomycin is a cooperative process. Binding of dpA tp actinomycin is partially cooperative and weaker than binding of dpG. The dimerization constant of actinomycin was also determined by the phase separation technique, and found in agreement with other values, including the results of kinetic measurements reported here.  相似文献   

17.
NMP kinases catalyse the phosphorylation of the canonical nucleotides to the corresponding diphosphates using ATP as a phosphate donor. Bacteriophage T4 deoxynucleotide kinase (DNK) is the only member of this family of enzymes that recognizes three structurally dissimilar nucleotides: dGMP, dTMP and 5-hydroxymethyl-dCMP while excluding dCMP and dAMP. The crystal structure of DNK with its substrate dGMP has been determined at 2.0 A resolution by single isomorphous replacement. The structure of the ternary complex with dGMP and ATP has been determined at 2.2 A resolution. The polypeptide chain of DNK is folded into two domains of equal size, one of which resembles the mononucleotide binding motif with the glycine-rich P-loop. The second domain, consisting of five alpha-helices, forms the NMP binding pocket. A hinge connection between the domains allows for large movements upon substrate binding which are not restricted by dimerization of the enzyme. The mechanism of active centre formation via domain closure is described. Comparison with other P-loop-containing proteins indicates an induced-fit mode of NTP binding. Protein-substrate interactions observed at the NMP and NTP sites provide the basis for understanding the principles of nucleotide discrimination.  相似文献   

18.
19.
We conclude from X-ray diffraction studies at low resolution (7 Å) that the binding of sugar and nucleotide substrates to dimeric yeast hexokinase BII crystals exhibits both negative co-operativity and positive allosteric co-operativity. Difference electron density maps show the positions of sugar and nucleotide binding sites and extensive substrate-induced structural changes in the protein. Sugar substrates and inhibitors bind in the deep cleft that divides each subunit into two lobes and nucleotide substrates bind nearby to one site per dimer, which lies between the subunits and on the molecular symmetry axis. Although the inhibitors o- and p-iodobenzoylglucosamine and o-toluoylglucosamine bind equally to both subunits, the degree of substitution of glucose or xylose is very different for the two subunits. The substrate analog β, γ-imido ATP shows only one strong binding site per dimer. This negative co-operativity in substrate binding may result from the heterologous or non-equivalent association of the two subunits (Anderson et al., 1974), which provides non-equivalent environments for the two chemically identical subunits.Further, there is a positive allosteric interaction between the sugar and nucleotide binding sites. Sugar binding is required for nucleotide binding at the intersubunit site and the binding of nucleotide modifies the binding of sugars. These positive heterotropic interactions appear to be mediated by extensive substrate-induced structural changes in the enzyme.  相似文献   

20.
Properties of caldesmon isolated from chicken gizzard.   总被引:5,自引:4,他引:1       下载免费PDF全文
Chicken gizzard smooth muscle contains two major calmodulin-binding proteins: caldesmon (11.1 microM; Mr 141 000) and myosin light-chain kinase (4.6 microM; Mr 136 000), both of which are associated with the contractile apparatus. The amino acid composition of caldesmon is distinct from that of myosin light-chain kinase and is characterized by a very high glutamic acid content (25.5%), high contents of lysine (13.6%) and arginine (10.3%), and a low aromatic amino acid content (2.4%). Caldesmon lacked myosin light-chain kinase and phosphatase activities and did not compete with either myosin light-chain kinase or cyclic nucleotide phosphodiesterase (both calmodulin-dependent enzymes) for available calmodulin, suggesting that calmodulin may have distinct binding sites for caldesmon on the one hand and myosin light-chain kinase and cyclic nucleotide phosphodiesterase on the other. Consistent with the lack of effect of caldesmon on myosin phosphorylation, caldesmon did not affect the assembly or disassembly of myosin filaments in vitro. As previously shown [Ngai & Walsh (1984) J. Biol. Chem. 259, 13656-13659], caldesmon can be reversibly phosphorylated. The phosphorylation and dephosphorylation of caldesmon were further characterized and the Ca2+/calmodulin-dependent caldesmon kinase was purified; kinase activity correlated with a protein of subunit Mr 93 000. Caldesmon was not a substrate of myosin light-chain kinase or phosphorylase kinase, both calmodulin-activated protein kinases.  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