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 15 毫秒
1.
Circular dichroism has been used to monitor the binding of pyridine nucleotide cofactors to enzyme-folate analog complexes of dihydrofolate reductase from Escherichia coli B (MB 1428). The enzyme binds one molar equivalent of many folate analogs and two molar equivalents of several pyridine nucleotide cofactors. The apo-enzyme has very low optical activity. The binding of folate analogs including folate, dihydrofolate, methotrexate, trimethoprim and pyrimethamine induce large Cotton effects. Pyridine nucleotides when bound to the enzyme-folate analog complexes also induce new optically active bands; all the effects being due to the first molar equivalent of cofactor bound. NADPH and NADP+ induce very similar bands when bound to the enzyme-methotrexate complex suggesting that the geometry of the complexes formed are very similar. The oxidized and reduced cofactor likewise have similar effects on the enzyme-folate complex. However, NADPH and NADP+ addition to both the enzyme-trimethoprim and enzyme-pyrimethamine complexes have significantly different effects on the circular dichroism spectra, suggesting that the inhibitors which are less homologous to the natural dihydrofolate substrate allow more conformational freedom in the enzyme-inhibitor-cofactor complex. In most cases the prior binding of the folate analog greatly increases the binding of the first molar equivalent of cofactor so that at concentrations of approx. 5-20 muM the binding appears stoichiometric. Pyrimethamine is an exception in that it apparently has no effect on the binding of NADPH to the enzyme.  相似文献   

2.
Elongation factor (EF)-1 beta, a 26 kDa protein, is the eukaryotic equivalent of bacterial EF-Ts, the nucleotide exchange factor in protein synthesis. EF-1 beta catalyzes the exchange of guanine nucleotides bound to EF-1 alpha; the latter protein is the eukaryotic equivalent of bacterial EF-Tu. Limited proteolytic cleavage studies on EF-1 beta lead to the following picture: the protein is composed of two domains, an aminoterminal and a carboxyterminal domain, connected to each other by a stretch of hydrophilic, charged amino acids situated in the middle of the molecule. The carboxyterminal domain supplies the catalytic site for the nucleotide exchange reaction, whereas the aminoterminal domain interacts with EF-1 gamma, the third component of elongation factor 1. The regulatory, serine phosphate residue, Ser-89, localized in the hydrophilic stretch of EF-1 beta, does not appear to be necessary for the basic exchange reaction. The fourth component of the high molecular weight elongation factor complex (EF-1H), named EF-1 delta or 28 K protein, is homologous to EF-1 beta and contains regions very similar to the carboxyterminal part. EF-1 delta was found to be active in the nucleotide exchange reaction.  相似文献   

3.
The binary complex of pig plasma gelsolin with Mg2+-G-actin in ATP and ADP   总被引:1,自引:0,他引:1  
H E Harris 《FEBS letters》1988,233(2):359-362
Pig plasma gelsolin combined with Mg-G-actin at less than 10(-8) M Ca2+ to yield a binary complex. Complexes formed from G-actin with bound ATP or ADP. They contained approx. 1 mol of non-exchangeable nucleotide per mol of actin. ATP hydrolysis was not coupled to binary complex formation, but ATP in the complex hydrolysed very slowly. The nucleotide in the binary complex behaved like one of the two nucleotide molecules in the ternary complex (two actin monomers to one gelsolin), but the actin-gelsolin interaction was weaker in the binary complex.  相似文献   

4.
Formation of the 5-fluorodeoxyuridylate-thymidylate synthetase binary complex generates a 19F nmr resonance 1.3–1.4 ppm to higher shielding from free ligand, probably as the result of rotation of the pyrimidine ring about the glycosyl bond. Addition of sodium dodecyl sulfate to the complex produces the spectrum of free ligand indicating that in contrast to the ternary complex of enzyme:nucleotide:cofactor, the binary complex does not contain a covalent bond linking the nucleotide to the enzyme. In the presence of a 2.5 molar excess of nucleotide, 1.55 moles were bound per mole of enzyme in Tris-Cl buffer. Under comparable conditions in sodium phosphate, 0.64 moles were bound, suggesting a specific buffer effect by phosphate.  相似文献   

5.
S.B. Marston  R.T. Tregear 《BBA》1974,333(3):581-584
The nucleotides bound to calcium-activated muscle fibres or myofibrils were separated and estimated. Most of the bound nucleotide was found to be ADP, as in relaxed muscle. The maximum ADP binding was not significantly altered by activation but the dissociation constant of the bound ADP was slightly reduced. These results are discussed in terms of the possibility of formation of a ternary ADP—myosin—actin complex.  相似文献   

6.
Prochniewicz E  Walseth TF  Thomas DD 《Biochemistry》2004,43(33):10642-10652
We have used optical spectroscopy (transient phosphorescence anisotropy, TPA, and fluorescence resonance energy transfer, FRET) to detect the effects of weakly bound myosin S1 on actin during the actomyosin ATPase cycle. The changes in actin were reported by (a) a phosphorescent probe (ErIA) attached to Cys 374 and (b) a FRET donor-acceptor pair, IAEDANS attached to Cys 374 and a nucleotide analogue (TNPADP) in the nucleotide-binding cleft. Strong interactions were detected in the absence of ATP, and weak interactions were detected in the presence of ATP or its slowly hydrolyzed analogue ATP-gamma-S, under conditions where a significant fraction of weakly bound acto-S1 complex was present and the rate of nucleotide hydrolysis was low enough to enable steady-state measurements. The results show that actin in the weakly bound complex with S1 assumes a new structural state in which (a) the actin filament has microsecond rotational dynamics intermediate between that of free actin and the strongly bound complex and (b) S1-induced changes are not propagated along the actin filament, in contrast to the highly cooperative changes due to the strongly bound complex. We propose that the transition on the acto-myosin interface from weak to strong binding is accompanied by transitions in the structural dynamics of actin parallel to transitions in the dynamics of interacting myosin heads.  相似文献   

7.
In order to investigate the flexibility of the ternary complex consisting of myosin subfragment-1 (S1), ADP, and orthovanadate (Vi), i.e., S1.ADP.Vi, the exchangeability of the bound ADP was examined. After isolation of the ternary complex of S1.ADP.Vi by gel filtration, 3'-O-(N-methylanthraniloyl)-ADP (Mant-ADP), a fluorescent analogue of ADP, was added at 0.5 degrees C. The added Mant-ADP was incorporated into the ternary complex very slowly by replacing the bound ADP. The nucleotide exchange occurred without regeneration of the ATPase activity of S1. Similarly, the ternary complex of S1.Mant-ADP.Vi prepared and isolated by gel filtration according to Hiratsuka (3, 4), was incubated with ADP (2.4 mM) at 4.5 degrees C. The nucleotide exchange of S1.Mant-ADP.Vi with ADP occurred in two phases with the apparent rates of 4.5 x 10(-4) s-1 (the fast phase) and 6.7 x 10(-6) s-1 (the slow phase). Biphasic exchange of the bound nucleotide was also observed with S1(A1) isozyme, indicating that the biphasic exchange did not correspond to two S1 isozymes. The apparent rates of the fast and the slow phases increased with the concentration of the added ADP, but they became saturated at an ADP concentration of the order of 2 mM, indicating that the nucleotide exchange reaction involves a step (or steps) which is insensitive to the concentration of free ADP in the solution. This step might be a reversible isomerization.  相似文献   

8.
Hyperphosphorylation of the Drosophila melanogaster origin recognition complex (DmORC) by cyclin dependent kinases (CDKs) allows nucleotide binding but inhibits the ATPase activity of Orc1, and ablates the ATP-dependent interaction of ORC with DNA. Here we present single particle electron microscopy (EM) studies of ORC bound to nucleotide in both the dephosphorylated and hyper-phosphorylated states. 3D image reconstructions show that nucleotide binding gives rise to an analogous conformation independent of phosphorylation state. At the intermediate resolution achieved in our studies, ATP promotes changes along the toroidal core of the complex with negligible differences contributed by phosphorylation. Thus, hyperphosphorylation of DmORC does not induce meso-scale rearrangement of the ORC structure. To better understand ORC's role in origin remodeling, we performed atomic force microscopy (AFM) studies that show the contour length of a 688bp linear DNA fragment shortens by the equivalent of approximately 130bp upon ORC binding. This data, coupled with previous studies that showed a linking number change in circular DNA upon ORC binding, suggests that ORC may wrap the DNA in a manner akin to DnaA. Based on existing data and our structures, we propose a subunit arrangement for the AAA+ and winged helix domains, and in addition, speculate on a path of the 133bp of DNA around the ORC complex.  相似文献   

9.
Recent crystal structures have revealed that regulatory subunit RIalpha of PKA undergoes a dramatic conformational change upon complex formation with the catalytic subunit. Molecular dynamics studies were initiated to elucidate the contributions of intrinsic conformational flexibility and interactions with the catalytic subunit in formation and stabilization of the complex. Simulations of a single RIalpha nucleotide binding domain (NBD), missing cAMP, showed that its C helix spontaneously occupies two distinct conformations: either packed against the nucleotide binding domain as in its cAMP bound structure, or extended into an intermediate form resembling that of the holoenzyme structure. C helix extension was not seen in a simulation of either RIalpha NBD. In a model complex containing both NBDs and the catalytic subunit, well-conserved residues at the interface between the NBDs in the cAMP bound form were found to stabilize the complex through contacts with the catalytic subunit. The model structure is consistent with available experimental data.  相似文献   

10.
Hsp110 is a nucleotide-activated exchange factor for Hsp70   总被引:1,自引:0,他引:1  
Hsp110 proteins constitute a subfamily of the Hsp70 chaperones and are potent nucleotide exchange factors (NEFs) for canonical Hsp70s of the eukaryotic cytosol. Here, we show that the NEF activity of the yeast Hsp110 homologue Sse1 itself is controlled by nucleotide. Nucleotide binding results in formation of a stabilized conformation of Sse1 that is required for association with the yeast Hsp70 Ssa1. The interaction triggers release of bound ADP from Ssa1, but nucleotide persists bound to Sse1 in the complex. Surprisingly, removal of this nucleotide does not affect the integrity of the complex. Instead, rebinding of ATP to the Hsp70 prompts the dissociation of the complex. Our data demonstrate that in contrast to previously characterized NEFs for Hsp70 chaperones, the NEF activity of Sse1 requires nucleotide binding and let us propose a new model for Hsp110 function.  相似文献   

11.
We have determined two different structures of PcrA DNA helicase complexed with the same single strand tailed DNA duplex, providing snapshots of different steps on the catalytic pathway. One of the structures is of a complex with a nonhydrolyzable analog of ATP and is thus a "substrate" complex. The other structure contains a bound sulphate ion that sits in a position equivalent to that occupied by the phosphate ion produced after ATP hydrolysis, thereby mimicking a "product" complex. In both complexes, the protein is monomeric. Large and distinct conformational changes occur on binding DNA and the nucleotide cofactor. Taken together, these structures provide evidence against an "active rolling" model for helicase action but are instead consistent with an "inchworm" mechanism.  相似文献   

12.
A molecular dynamics investigation and coarse-grained analysis of inactivated actin-related protein (Arp) 2/3 complex is presented. It was found that the nucleotide binding site within Arp3 remained in a closed position with bound ATP or ADP, but opened when simulation with no nucleotide was performed. In contrast, simulation of the isolated Arp3 subunit with bound ATP, showed a fast opening of the nucleotide binding cleft. A homology model for the missing subdomains 1 and 2 of Arp2 was constructed, and it was also found that the Arp2 binding cleft remained closed with bound nucleotide. Within the nucleotide binding cleft a distinct opening and closing period of 10 ns was observed in many of the simulations of Arp2/3 as well as isolated Arp3. Substitution studies were employed, and several alanine substitutions were found to induce a partial opening of the ATP binding cleft in Arp3 and Arp2, whereas only a single substitution was found to induce opening of the ADP binding cleft. It was also found that the nucleotide type did not cause a substantial change on interfacial contacts between Arp3 and the ArpC2, ArpC3 and ArpC4 subunits. Nucleotide-free Arp3 had generally less stable contacts, but the overall contact architecture was constant. Finally, nucleotide-dependent coarse-grained models for Arp3 are developed that serve to further highlight the structural differences induced in Arp3 by nucleotide hydrolysis.  相似文献   

13.
Fast, accurate nucleotide incorporation by polymerases facilitates expression and maintenance of genomes. Many polymerases use conformational dynamics of a conserved α helix to permit efficient nucleotide addition only when the correct nucleotide substrate is bound. This α helix is missing in structures of RNA-dependent RNA polymerases (RdRps) and RTs. Here, we use solution-state nuclear magnetic resonance to demonstrate that the conformation of conserved structural motif D of an RdRp is?linked to the nature (correct versus incorrect) of the bound nucleotide and the protonation state of a conserved, motif-D lysine. Structural data also reveal the inability of motif D to achieve its optimal conformation after incorporation of an incorrect nucleotide. Functional data are consistent with the conformational change of motif D becoming rate limiting during and after nucleotide misincorporation. We conclude that motif D of RdRps and, by inference, RTs is the functional equivalent to the fidelity helix of other polymerases.  相似文献   

14.
The H-ras gene product p21H has been mutated at Phe-28, which makes a hydrophobic interaction with the guanine base of bound GDP/GTP. The mutation Phe-28----Leu drastically increases nucleotide dissociation rates without affecting association rates. This is due to a perturbed binding of base, alpha- and beta-phosphate, and Mg2+, as evidenced from 31P NMR and fluorescence measurements. The region around the gamma-phosphate appears normal. The affinity of Mg2+ for both the di- and the triphosphate conformation of the mutant was also measured by fluorescence. The association constant is 3.5 x 10(7) M-1 for the Gpp(NH)p complex, 500 times higher than for the GDP form. The mutation does not change appreciably the intrinsic or the GTPase activating protein (GAP)-stimulated GTPase. The mutated protein induces neurite differentiation however when pressure-loaded into PC12 cells, which is equivalent to transformation of NIH 3T3 cells. This shows that p21 (F28L) is converted to the GDP bound form by GAP but is transforming because the high dissociation rate for nucleotides leads to a protein predominantly in the active GTP bound form.  相似文献   

15.
During mismatch repair ATP binding and hydrolysis activities by the MutS family proteins are important for both mismatch recognition and for transducing mismatch recognition signals to downstream repair factors. Despite intensive efforts, a MutS.ATP.DNA complex has eluded crystallographic analysis. Searching for ATP analogs that strongly bound to Thermus aquaticus (Taq) MutS, we found that ADP.beryllium fluoride (ABF), acted as a strong inhibitor of several MutS family ATPases. Furthermore, ABF promoted the formation of a ternary complex containing the Saccharomyces cerevisiae MSH2.MSH6 and MLH1.PMS1 proteins bound to mismatch DNA but did not promote dissociation of MSH2.MSH6 from mismatch DNA. Crystallographic analysis of the Taq MutS.DNA.ABF complex indicated that although this complex was very similar to that of MutS.DNA.ADP, both ADP.Mg(2+) moieties in the MutS. DNA.ADP structure were replaced by ABF. Furthermore, a disordered region near the ATP-binding pocket in the MutS B subunit became traceable, whereas the equivalent region in the A subunit that interacts with the mismatched nucleotide remained disordered. Finally, the DNA binding domains of MutS together with the mismatched DNA were shifted upon binding of ABF. We hypothesize that the presence of ABF is communicated between the two MutS subunits through the contact between the ordered loop and Domain III in addition to the intra-subunit helical lever arm that links the ATPase and DNA binding domains.  相似文献   

16.
A palmitoyl CoA-protein complex was isolated from bovine heart mitochondria and purified to homogeneity. The elution profile of the [14C]palmitoyl CoA bound protein from a hydroxyapatite column was identical to that seen when [3H]carboxyatractylate was used as the bound ligand. A sample of the palmitoyl CoA-protein complex from a peak fraction of the column appeared to be homogeneous by sodium dodecyl sulfate gel electrophoresis. The mobility of the protein bound with palmitoyl CoA was identical to the one bound with carboxyatractylate and the molecular weight was estimated to be 30,000 daltons. Compared to the stable palmitoyl CoA-protein complex, purification of the unliganded carrier from mitochondria at 22°C resulted in a disaggregated protein. These physical characteristics of the palmitoyl CoA-protein complex correspond to those identified for the ADPATP carrier. The results further confirm the specificity of the fatty acyl CoA ligand for the adenine nucleotide translocase and support the concept that it may be a physiological modulator of adenine nucleotide translocation.  相似文献   

17.
S Beharry  P D Bragg 《Biochemistry》1992,31(46):11472-11476
Escherichia coli F1-ATPase contained 3 mol of tightly-bound adenine nucleotide/mol enzyme. A further 3 mol could be loaded by incubation of the enzyme with ATP. The unloaded enzyme was designated as a F1[2,1] type on the basis of the ability of GTP to displace 1 mol of adenine nucleotide/mol of F1 [Kironde, F.A.S., & Cross, R.L. (1986) J. Biol. Chem. 261, 12544-12549]. The loaded enzyme was designated F1[3,3] since GTP could displace 3 of the 6 mol of bound adenine nucleotide/mol of F1. Incubation of F1[2,1], F1[2,0], and F1[3,0] with phosphate in the presence of 30% (v/v) dimethyl sulfoxide led to the synthesis of ATP from endogenous bound ADP. Hydrolysis of newly synthesized ATP occurred on transfer of the F1 from 30% (v/v) dimethyl sulfoxide to an entirely aqueous medium. Thus, synthesis and hydrolysis of ATP can occur at GTP-nonchaseable adenine nucleotide binding sites, and these sites in dimethyl sulfoxide are not necessarily equivalent to noncatalytic sites.  相似文献   

18.
Succinyl-CoA synthetase (SCS) catalyzes the following reversible reaction via a phosphorylated histidine intermediate (His 246alpha): succinyl-CoA + P(i) + NDP <--> succinate + CoA + NTP (N denotes adenosine or guanosine). To determine the structure of the enzyme with nucleotide bound, crystals of phosphorylated Escherichia coli SCS were soaked in successive experiments adopting progressive strategies. In the first experiment, 1 mM ADP (>15 x K(d)) was added; Mg(2+) ions were omitted to preclude the formation of an insoluble precipitate with the phosphate and ammonium ions. X-ray crystallography revealed that the enzyme was dephosphorylated, but the nucleotide did not remain bound to the enzyme (R(working) = 17.2%, R(free) = 22.8% for data to 2.9 A resolution). Catalysis requires Mg(2+) ions; hence, the "true" nucleotide substrate is probably an ADP-Mg(2+) complex. In the successful experiment, the phosphate buffer was exchanged with MOPS, the concentration of sulfate ions was lowered, and the concentrations of ADP and Mg(2+) ions were increased to 10.5 and 50 mM, respectively. X-ray diffraction data revealed an ADP-Mg(2+) complex bound in the ATP-grasp fold of the N-terminal domain of each beta-subunit (R(working) = 19.1%, R(free) = 24.7% for data to 3.3 A resolution). We describe the specific interactions of the nucleotide-Mg(2+) complex with SCS, compare these results with those for other proteins containing the ATP-grasp fold, and present a hypothetical model of the histidine-containing loop in the "down" position where it can interact with the nucleotide approximately 35 A from where His 246alpha is seen in both phosphorylated and dephosphorylated SCS.  相似文献   

19.
C L Berger  D D Thomas 《Biochemistry》1991,30(46):11036-11045
We have used saturation-transfer electron paramagnetic resonance (ST-EPR) to detect the microsecond rotational motions of spin-labeled myosin subfragment one (MSL-S1) bound to actin in the presence of the ATP analogues AMPPNP (5'-adenylylimido diphosphate) and ATP gamma S [adenosine 5'-O-(3-thiotriphosphate)], which are believed to trap myosin in strongly and weakly bound intermediate states of the actomyosin ATPase cycle, respectively. Sedimentation binding measurements were used to determine the fraction of myosin heads bound to actin under ST-EPR conditions and the fraction of heads containing bound nucleotide. ST-EPR spectra were then corrected to obtain the spectrum corresponding to the ternary complex (actin.MSL-S1.nucleotide). The ST-EPR spectrum of MSL-S1.AMPPNP bound to actin is identical to that obtained in the absence of nucleotide (rigor complex), indicating no rotational motion of MSL-S1 relative to actin on the microsecond time scale. However, MSL-S1-ATP gamma S bound to actin is rotationally mobile, with an effective rotational correlation time (tau r) of 17 +/- 2 microseconds. This motion is similar to that observed previously for actin-bound MSL-S1 during the steady-state hydrolysis of ATP [Berger et al. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 8753-8757]. We conclude that, in solution, the weakly bound actin-attached states of the myosin ATPase cycle undergo microsecond rotational motions, while the strongly bound intermediates do not, and that these motions are likely to be involved in the molecular mechanism of muscle contraction.  相似文献   

20.
Measurements of the relaxation rate of water protons (PRR) have been used to study the interaction of yeast phosphoglycerate kinase with the manganous complexes of a number of nucleotides. The results indicate that phosphoglycerate kinase belongs to the same class of enzymes as creatine kinase, adenylate kinase, formyltetrahydrofolate synthetase, and arginine kinase, with maximal binding of metal ion to tne enzyme in the presence of the nucleotide substrate. However, an analysis of titration curves for a number of nucleoside diphosphates (ADP, IDP, GDP) showed that there is a substantial synergism in binding of the metal ion and nucleotide to the enzyme in the ternary complex. The metal-substrate binds to the enzyme approximately two orders of magnitude more tightly than the free nucleotide; Other evidence for an atypical binding scheme for Mn(II)-nucleoside diphosphates was obtained by electron paramagnetic resonance (EPR) studies; the EPR spectrum for the bound Mn(II) in the enzyme-MnADP complex differed substantially from those obtained for other kinases. An identical EPR spectrum is observed with the MnADP complex with the rabbit muscle enzyme as with the yeast enzyme. In contrast, the dissociation constant for the enzyme-MnATP complex is approximately fourfold lower than that for enzyme-ATP, and there are no substantial changes in the electron paramagnetic resonance spectrum of MnATP2- when the complex is bound to phosphoglycerate kinase. A small but significant change in the PRR of water is observed on addition of 3-phosphoglycerate (but not 2-phosphoglycerate) to the MnADP-enzyme complex. However, addition of 3-phosphoglycerate to enzyme-MnADP did not influence the EPR spectrum of the enzyme-bound Mn(II).  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