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.
Escherichia coli glycerol kinase (GK) displays "half-of-the-sites" reactivity toward ATP and allosteric regulation by fructose 1, 6-bisphosphate (FBP), which has been shown to promote dimer-tetramer assembly and to inhibit only tetramers. To probe the role of tetramer assembly, a mutation (Ser58-->Trp) was designed to sterically block formation of the dimer-dimer interface near the FBP binding site [Ormo, M., Bystrom, C., and Remington, S. J. (1998) Biochemistry 37, 16565-16572]. The substitution did not substantially change the Michaelis constants or alter allosteric regulation of GK by a second effector, the phosphocarrier protein IIAGlc; however, it eliminated FBP inhibition. Crystal structures of GK in complex with different nontransferable ATP analogues and glycerol revealed an asymmetric dimer with one subunit adopting an open conformation and the other adopting the closed conformation found in previously determined structures. The conformational difference is produced by a approximately 6.0 degrees rigid-body rotation of the N-terminal domain with respect to the C-terminal domain, similar to that observed for hexokinase and actin, members of the same ATPase superfamily. Two of the ATP analogues bound in nonproductive conformations in both subunits. However, beta, gamma-difluoromethyleneadenosine 5'-triphosphate (AMP-PCF2P), a potent inhibitor of GK, bound nonproductively in the closed subunit and in a putative productive conformation in the open subunit, with the gamma-phosphate placed for in-line transfer to glycerol. This asymmetry is consistent with "half-of-the-sites" reactivity and suggests that the inhibition of GK by FBP is due to restriction of domain motion.  相似文献   

2.
We analysed by analytical ultracentrifugation and fluorescence anisotropy the binding of p53 truncation mutants to sequence-specific DNA. The synthetic 30 base-pair DNA oligomers contained the 20 base-pair recognition elements for p53, consisting of four sites of five base-pairs per p53 monomer. We found that the binding at low ionic strengths was obscured by artifacts of non-specific binding and so made measurements at higher ionic strengths. Analytical ultracentrifugation of the construct p53CT (residues 94-360, containing the DNA-binding core and tetramerization domains) gave a dissociation constant of approximately 3 microM for its dimer-tetramer equilibrium, similar to that of full-length protein. Analytical ultracentrifugation and fluorescence anisotropy showed that p53CT formed a complex with the DNA constructs with 2:1 stoichiometry (dimer:DNA). The binding of p53CT (1-100 nm range) to DNA was highly cooperative, with a Hill coefficient of 1.8 (dimer:DNA). The dimeric L344A mutant of p53CT has impaired tetramerization. It bound to full-length DNA p53 recognition sequence, but with sixfold less affinity than wild-type protein. It did not form a detectable complex with a 30-mer DNA construct containing two specific five base-pair sites and two random sites, emphasizing the high co-operativity of the binding. The fundamental active unit of p53 appears to be the tetramer, which is induced by DNA binding, although it is a dimer at low concentrations.  相似文献   

3.
J K de Riel  H Paulus 《Biochemistry》1978,17(24):5146-5150
The mechanism of desensitization of glycerol kinase to allosteric inhibition by fructose 1,6-bisphosphate caused by salt, urea, and high pH has been examined in the light of the model proposed in an earlier paper [de Riel, J. K., and Paulus H. (1978), Biochemistry 17] relating subunit dissociation and ligand binding. KCl (0.4 M) causes a tenfold decrease in the affinity of tetrameric glycerol kinase for fructose, 1,6-bisphosphate but has no significant effect on the dissociation process itself. Urea (2 M) causes a large increase in the equilibrium constant for the dissociation of the glycerol kinase tetramer to dimer but has no effect on the affinity of the tetramer for the allosteric inhibitor. High pH (9--10) has only a small effect on the subunit dissociation constant but greatly reduces the rates of subunit association and dissociation. Desensitization of glycerol kinase to allosteric inhibition can thus occur by three different mechanisms, two of which are directly related to the polysteric nature of the enzyme.  相似文献   

4.
C A Royer  G Weber  T J Daly  K S Matthews 《Biochemistry》1986,25(25):8308-8315
Dissociation of lac repressor tetramer by high hydrostatic pressures was monitored with intrinsic tryptophan fluorescence. With the assumption of complete dissociation to monomer, tryptophan polarization data gave delta V a approximately 170 mL/mol and the concentration for 50% tetramer dissociation, C1/2, was 3.8 X 10(-8) M. Upon addition of inducer, the calculated delta V a increased to approximately 220 mL/mol and the C1/2 decreased to approximately 1 X 10(-8) M, a free energy difference of approximately 0.7 kcal. These results indicate a modest stabilization of the tetramer by the presence of inducer. Monitoring the average energy of tryptophan emission demonstrated that tetramer dissociation takes place over the same range of pressures as evidenced by the polarization data and IPTG dissociation can be more or less superimposed upon tetramer dissociation depending upon the ligand concentration used. Although the two transitions cannot be separated entirely, the delta V a for the region of the pressure dependence dominated by ligand dissociation was 69 mL/mol, an unexpectedly large value. For tetramer modified with methyl methanethiosulfonate, subunit dissociation was shifted to much higher pressures and IPTG dissociation did not occur. The delta V a for subunit association was calculated as approximately 160 mL/mol, and the C1/2 was 3.5 X 10(-9) M. Interactions at the subunit interface of the modified protein are apparently stronger than in the unmodified protein. The absence of inducer dissociation from the MMTS-modified tetramer by the application of high hydrostatic pressure suggests that the volume change for inducer binding to the modified protein is much smaller than that observed for the unmodified repressor.  相似文献   

5.
Kinetic theory of dissociating enzyme systems has been applied to a study of the dimer-tetramer interconversion of glycogen phosphorylase a. All kinetic constants for the dissociating-associating reaction of phosphorylase a have been determined. The results indicate that (a) the presence of glucose-1-phosphate has no influence on either the rate of dissociation or the rate of association, and hence does not shift the dimer-tetramer equilibrium of phosphorylase a; (b) the binding og glycogen to the enzyme decreases the association rate of the dimer to form the tetramer, but has no effect on the dissociation rate of the tetramer; (c) both the dimeric and tetrameric form of phosphorylase a can bind glycogen, but the tetrameric form has a lower affinity for glycogen and is catalytically inactive.  相似文献   

6.
The reversible inactivation and dissociation of the allosteric phosphofructokinase from Escherichia coli has been studied in relatively mild conditions, i.e., in the presence of the chaotropic agent KSCN. At moderate KSCN concentration, the loss of enzymatic activity involves two separated phases: first, a rapid dissociation of part of the tetramer into dimers, second, a slower displacement of the dimer-tetramer equilibrium upon further dissociation of the dimer into monomers. These two reactions can no longer be distinguished above 0.3 M KSCN since complete inactivation occurs in a single reaction. Different changes are observed for the fluorescence and the activity of the enzyme in KSCN: the fluorescence is not affected by the dissociation into dimers which is responsible for inactivation. The decrease in fluorescence reflects the change in environment of the unique tryptophan residue, Trp 311, during the dimer to monomer dissociation. This residue belongs to the interface containing the regulatory site, and its native fluorescence indicates that this interface is still present in the dimer. The substrate fructose 6-phosphate protects phosphofructokinase from inactivation by binding to the tetramer and prevents its dissociation into dimers. The presence of phosphoenolpyruvate prevents the slow dissociation of the dimer into monomers, which shows the ability of the dimer to bind the inhibitor. Two successive processes can be observed during reassociation of the protein upon KSCN dilution. First, a fast reaction (k1 = 2 x 10(5) M-1.s-1) is accompanied by a fluorescence increase and results in the formation of the dimeric species.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
The unfolding and dissociation of the tetrameric enzyme fructose-1,6-bisphosphatase from pig kidney by guanidine hydrochloride have been investigated at equilibrium by monitoring enzyme activity, ANS binding, intrinsic (tyrosine) protein fluorescence, exposure of thiol groups, fluorescence of extrinsic probes (AEDANS, MIANS), and size-exclusion chromatography. The unfolding is a multistate process involving as the first intermediate a catalytically inactive tetramer. The evidence that indicates the existence of this intermediate is as follows: (1) the loss of enzymatic activity and the concomitant increase of ANS binding, at low concentrations of Gdn.HCl (midpoint at 0.75 M), are both protein concentration independent, and (2) the enzyme remains in a tetrameric state at 0.9 M Gdn.HCl as shown by size-exclusion chromatography. At slightly higher Gdn.HCl concentrations the inactive tetramer dissociates to a compact dimer which is prone to aggregate. Further evidence for dissociation of tetramers to dimers and of dimers to monomers comes from the concentration dependence of AEDANS-labeled enzyme anisotropy data. Above 2.3 M Gdn.HCl the change of AEDANS anisotropy is concentration independent, indicative of monomer unfolding, which also is detected by a red shift of MIANS-labeled enzyme emission. At Gdn.HCl concentrations higher than 3.0 M, the protein elutes from the size-exclusion column as a single peak, with a retention volume smaller than that of the native protein, corresponding to the completely unfolded monomer. In the presence of its cofactor Mg(2+), the denaturated enzyme could be successfully reconstituted into the active enzyme with a yield of approximately 70-90%. Refolding kinetic data indicate that rapid refolding and reassociation of the monomers into a nativelike tetramer and reactivation of the tetramer are sequential events, the latter involving slow and small conformational rearrangements in the refolded enzyme.  相似文献   

8.
Fructose 1,6-bisphosphate (FBP) stimulates the reaction of Bacillus stearothermophilus acetate kinase (AK). FBP changes the reaction curve for ATP from a sigmoidal type to a Michaelis-Menten one. The binding of FBP to AK was studied by an equilibrium dialysis method and by measuring changes in fluorescence. The extent of binding of FBP to the enzyme paralleled its activation. In addition, the binding constant for FBP increased in the presence of substrate, ATP. These results suggest that FBP is an allosteric activator of B. stearothermophilus AK. Only two moles of FBP bound to this tetrameric enzyme. No cooperativity was found for the binding of FBP. These observations support the previous conclusion, that a set of two subunits in the tetramer is a unit of the enzymatic function. A model is presented to interpret the sigmoidal kinetics for ATP, the absence of cooperativity for FBP binding, and the allosteric activation by FBP of this enzyme. The kinetic properties of the enzyme can be explained quantitatively by this model.  相似文献   

9.
Steady-state fluorescence anisotropy measurements can be used to detect fluorescence resonance energy transfer (FRET) between identical fluorophores (homo-FRET). However, the contribution of homo-FRET to the steady-state anisotropy must be discerned from those due to the orientational distribution and rotational diffusion, which so far has required photobleaching controls, largely precluding dynamic measurements in live cells. We describe a variation of steady-state anisotropy microscopy in which the contribution of homo-FRET is dynamically isolated from the total anisotropy by exploiting the loss of energy transfer that occurs at red-edge excitation. Excitation of enhanced green fluorescent protein (EGFP) at the red-edge of its absorption band shows the shift in the emission spectrum compared to main-band excitation that is characteristic for photo-selection of static low energy S(0)-S(1) transitions that fail to exhibit FRET. An experimental setup for steady-state fluorescent anisotropy microscopy is described that can be used to acquire anisotropy images in live cells at main-band and red-edge excitation of EGFP. We demonstrate in live cells homo-FRET suppression of protein fusion constructs that consist of two and three EGFP molecules connected by short linkers. This methodology represents a novel approach for the dynamic measurement of homo-FRET in live cells that will be of utility in the biological sciences to detect oligomerization and concentration dependent interactions between identically labeled molecules.  相似文献   

10.
High hydrostatic pressure coupled with fluorescence polarization has been used to investigate protein subunit interactions and protein-operator association in lac repressor labeled with a long-lived fluorescent probe. On the basis of observation of a concentration-dependent sigmoidal decrease in the dansyl fluorescence polarization, we conclude that application of high hydrostatic pressure results in dissociation of the lac repressor tetramer. The 2-fold decrease in the rotational relaxation time and the high-pressure plateau are consistent with a tetramer to dimer transition. The volume change for tetramer dissociation to dimer is -82 +/- 5 mL/mol. The dissociation constant calculated from the data taken at 4.5 degrees C is 4.3 +/- 1.3 nM. The tetramer dissociation constant increases by a factor of 3 when the temperature is raised from 4.5 to 21 degrees C. A very small effect of inducer binding on the subunit dissociation is observed at 4.5 degrees C; the Kd increases from 4.5 to 7.1 nM. At 21 degrees C, however, inducer binding stabilizes the tetramer by approximately 0.8 kcal/mol. Pressure-induced monomer formation is indicated by the curves obtained upon raising the pH to 9.2. The addition of IPTG shifts the pressure transition to only slightly higher pressures at this pH, indicating that the stabilization of the tetramer by inducer is not as marked as that observed at pH 7.1. From the decrease in the polarization of the dansyl repressor-operator complexes, we also conclude that the application of pressure results their dissociation and that the volume change is large in absolute value (approximately 200 mL/mol). The lac repressor-operator complex is more readily dissociated upon the application of pressure than the tetramer alone, indicating that operator binding destabilizes the lac repressor tetramer.  相似文献   

11.
Deoxyhemoglobin tetramers dissociate into dimers very slowly, with half-times on the order of several hours. It is demonstrated that absorbance changes in the Soret region which accompany this dissociation and persist upon binding of haptoglobin 1-1 to the dissociated dimers can be used for accurate kinetic determinations over the necessarily long periods required for study. This method of study for the slow reactions depends upon long-term spectral integrity of the reaction mixtures and upon accurate measurement. The variation in rate constants determined by this procedure has been correlated with variations in structural constraints at the dimer-dimer contact region. In the presence of 2,3-diphosphoglycerate the rate constant is decreased, consistent with the role of this effector in binding to both beta chains and stabilizing the constrained deoxy tetramer against dissociation into alphabeta dimers. With hemoglobin specifically modified (des-Arg-141alpha) to eliminate half the constraining salt links within the dimer-dimer contact region, the dissociation rate is increased by approximately three orders of magnitude. In hemoglobin S where the amino acid substitution is not directly in the intersubunit contact region of interest, the dissociation rate is found to be approximately the same as that for hemoglobin A. Combination of the dissociation rate constants determined by haptoglobin binding with stopped-flow determinations of the rate constant for reassociation of dissociated dimers provides an estimate of the equilibrium constant, 0K2, for the deoxyhemoglobin dimer-tetramer equilibrium. This estimate is independent of any assumptions regarding other energetic quantities, and yields a value of 2.54 +/- 0.7 X 10(10)M-1 (heme) in 0.1 M Tris-HCl, 0.1 M NaCl, and 1 mM EDTA, pH 7.4, 21.5 degrees C. Thus the intersubunit contact energy is -14.0 +/- 0.2 kcal/mol of heme. The stabilization energy between deoxy and oxy tetramers is found to be approximately 6.4 kcal/mol, under these conditions.  相似文献   

12.
13.
We have studied the effect of organic co-solvents (monohydric alcohols and formamide) on the dimer-tetramer equilibrium of human haemoglobin by measuring the dependence of oxygen affinity upon haemoglobin concentration over a 100-fold concentration range and analysing the data with a modified Monod-Wyman-Changeux model in which the equilibrium (tetramer) in equilibrium with (dimer) + (dimer) was taken into account. This procedure enabled us to obtain the dimer-tetramer equilibrium constant and to find its dependence upon co-solvent concentration. Then by following the procedure already reported for the tense----relaxed state transition of haemoglobin, we separated the co-solvent effects into bulk electrostatic and non-bulk electrostatic (hydrophobic) contributions. We believe that our results demonstrate that during haemoglobin dissociation, just as we have already shown for the tense to relaxed state transition, both charged groups and hydrophobic surfaces became exposed to the solvent.  相似文献   

14.
The crystal structure of an Escherichia coli glycerol kinase mutant Gly230 --> Asp (GKG230D) was determined to 2.0 A resolution using a microfluidics based crystallization platform. The crystallization strategy involved a suite of microfluidic devices that characterized the solubility trends of GKG230D, performed nanoliter volume free interface diffusion crystallization experiments, and produced diffraction-quality crystals for in situ data collection. GKG230D displays increased enzymatic activity and decreased allosteric regulation by the glycolytic pathway intermediate fructose 1,6-bisphosphate (FBP) compared to wild-type GK (GKWT). Structural analysis revealed that the decreased allosteric regulation is a result of the altered FBP binding loop conformations in GKG230D that interfere with the wild-type FBP binding site. The altered FBP binding loop conformations in GKG230D are supported through a series of intramolecular loop interactions. The appearance of Asp230 in the FBP binding loops also repositions the wild-type FBP binding residues away from the FBP binding site. Light scattering analysis confirmed GKG230D is a dimer and is resistant to tetramer formation in the presence of FBP, whereas GKWT dimers are converted into putatively inactive tetramers in the presence of FBP. GKG230D also provides the first structural evidence for multiple GK monomer conformations in the presence of glycerol and in the absence of a nucleotide substrate and verifies that glycerol binding is not responsible for locking GK into the closed conformation necessary for GK activity.  相似文献   

15.
The fructose-1,6-bis(phosphate) aldolase isologous tetramer tightly associates through two different subunit interfaces defined by its 222 symmetry. Both single- and double-interfacial mutant aldolases have a destabilized quaternary structure, but there is little effect on the catalytic activity. These enzymes are however thermolabile. This study demonstrates the temperature-dependent dissociation of the mutant enzymes and determines the dissociation free energies of both mutant and native aldolase. Subunit dissociation is measured by sedimentation equilibrium in the analytical ultracentrifuge. At 25 degrees C the tetramer-dimer dissociation constants for each single-mutant enzyme are similar, about 10(-6) M. For the double-mutant enzyme, sedimentation velocity experiments on sucrose density gradients support a tetramer-monomer equilibrium. Furthermore, sedimentation equilibrium experiments determined a dissociation constant of 10(-15) M3 for the double-mutant enzyme. By the same methods the upper limit for the dissociation constant of wild-type aldolase A is approximately 10(-28) M3, which indicates an extremely stable tetramer. The thermodynamic values describing monomer-tetramer and dimer-tetramer equilibria are analyzed with regard to possible cooperative interaction between the two subunit interfaces.  相似文献   

16.
The MutS DNA mismatch protein recognizes heteroduplex DNAs containing mispaired or unpaired bases. We have examined the oligomerization of a MutS protein from Thermus aquaticus that binds to heteroduplex DNAs at elevated temperatures. Analytical gel filtration, cross-linking of MutS protein with disuccinimidyl suberate, light scattering, and matrix-assisted laser desorption/ionization time-of-flight mass spectrometry establish that the Taq protein is largely a dimer in free solution. Analytical equilibrium sedimentation showed that the oligomerization of Taq MutS involves a dimer-tetramer equilibrium in which dimer predominates at concentrations below 10 microM. The DeltaG(0)(2-4) for the dimer to tetramer transition is approximately -6.9 +/- 0.1 kcal/mol of tetramer. Analytical gel filtration of native complexes and gel mobility shift assays of an maltose-binding protein-MutS fusion protein bound to a short, 37-base pair heteroduplex DNA reveal that the protein binds to DNA as a dimer with no change in oligomerization upon DNA binding.  相似文献   

17.
The role of phospholipid in the binding of coenzyme, NAD(H), to 3-hydroxybutyrate dehydrogenase, a lipid-requiring membrane enzyme, has been studied with the ultrafiltration binding method, which we optimized to quantitate weak ligand binding (KD in the range 10-100 microM). 3-Hydroxybutyrate dehydrogenase has a specific requirement of phosphatidylcholine (PC) for optimal function and is a tetramer quantitated both for the apodehydrogenase, which is devoid of phospholipid, and for the enzyme reconstituted into phospholipid vesicles in either the presence or absence of PC. We find that (i) the stoichiometry for NADH and NAD binding is 0.5 mol/mol of enzyme monomer (2 mol/mol of tetramer); (ii) the dissociation constant for NADH binding is essentially the same for the enzyme reconstituted into the mixture of mitochondrial phospholipids (MPL) (KD = 15 +/- 3 microM) or into dioleoyl-PC (KD = 12 +/- 3 microM); (iii) the binding of NAD+ to the enzyme-MPL complex is more than an order of magnitude weaker than NADH binding (KD approximately 200 microM versus 15 microM) but can be enhanced by formation of a ternary complex with either 2-methylmalonate (apparent KD = 1.1 +/- 0.2 microM) or sulfite to form the NAD-SO3- adduct (KD = 0.5 +/- 0.1 microM); (iv) the binding stoichiometry for NADH is the same (0.5 mol/mol) for binary (NADH alone) and ternary complexes (NADH plus monomethyl malonate); (v) binding of NAD+ and NADH together totals 0.5 mol of NAD(H)/mol of enzyme monomer, i.e., two nucleotide binding sites per enzyme tetramer; and (vi) the binding of nucleotide to the enzyme reconstituted with phospholipid devoid of PC is weak, being detected only for the NAD+ plus 2-methylmalonate ternary complex (apparent KD approximately 50 microM or approximately 50-fold weaker binding than that for the same complex in the presence of PC). The binding of NADH by equilibrium dialysis or of spin-labeled analogues of NAD+ by EPR spectroscopy gave complementary results, indicating that the ultrafiltration studies approximated equilibrium conditions. In addition to specific binding of NAD(H) to 3-hydroxybutyrate dehydrogenase, we find significant binding of NAD(H) to phospholipid vesicles. An important new finding is that the nucleotide binding site is present in 3-hydroxybutyrate dehydrogenase in the absence of activating phospholipid since (a) NAD+, as the ternary complex with 2-methylmalonate, binds to the enzyme reconstituted with phospholipid devoid of PC and (b) the apodehydrogenase, devoid of phospholipid, binds NADH or NAD-SO3- weakly (half-maximal binding at approximately 75 microM NAD-SO3- and somewhat weaker binding for NADH).(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
The Escherichia coli single strand binding (SSB) protein is an essential protein required for DNA replication and involved in recombination and a number of repair processes. It is a stable homotetramer in solution; however the ssb-1 mutation (His-55 to Tyr) destabilizes the tetramer with respect to monomers and this defect seems to explain the observed phenotype (Williams, K. R., Murphy, J. B., and Chase, J. W. (1984) J. Biol. Chem. 259, 11804-11811). We report a quantitative study of the SSB-1 monomer-tetramer equilibrium in vitro as a function of temperature, pH, NaCl, MgCl2, urea, and guanidine hydrochloride concentrations. The self-assembly equilibrium was monitored by the increase in intrinsic protein fluorescence anisotropy accompanying the formation of the tetramer. The experimental isotherms indicate that SSB-1 dimers are not highly populated at equilibrium, hence the formation of the tetramer is well-described as a one-step association of four monomers. At 25 degrees C, pH 8.1, the monomer concentration for 50% tetramer dissociation is (MT)1/2 = 0.87 microM, corresponding to a monomer-tetramer equilibrium constant, KT = 3 +/- 1 x 10(18) M-3. The tetramerization constant, KT, is highly dependent upon temperature and pH, with delta H0 = -51 +/- 7 kcal/mol (pH 8.1) and delta H0 = -37 +/- 5 kcal/mol (pH 6.9). There is no effect of NaCl on the monomer-tetramer association in the range from 0.20 to 1.0 M; however, MgCl2 decreases the stability of the SSB-1 tetramer. In the presence of high concentrations of the single-stranded oligonucleotide, dT(pT)15, the tetramerization constant is slightly increased indicating that binding of the oligonucleotide to the SSB-1 monomer promotes the assembly process, although not dramatically. The large negative delta H0 that is associated with formation of the tetramer provides a likely explanation for the temperature sensitivity of the ssb-1 mutation.  相似文献   

19.
D W Pettigrew  G J Yu  Y Liu 《Biochemistry》1990,29(37):8620-8627
Substrate binding to Escherichia coli glycerol kinase (EC 2.7.1.30; ATP-glycerol 3-phosphotransferase) was investigated by using both kinetics and binding methods. Initial-velocity studies in both reaction directions show a sequential kinetic mechanism with apparent substrate activation by ATP and substrate inhibition by ADP. In addition, the Michaelis constants differ greatly from the substrate dissociation constants. Results of product inhibition studies and dead-end inhibition studies using 5'-adenylyl imidodiphosphate show the enzyme has a random kinetic mechanism, which is consistent with the observed formation of binary complexes with all the substrates and the glycerol-independent MgATPase activity of the enzyme. Dissociation constants for substrate binding determined by using ligand protection from inactivation by N-ethylmaleimide agree with those estimated from the initial-velocity studies. Determinations of substrate binding stoichiometry by equilibrium dialysis show half-of-the-sites binding for ATP, ADP, and glycerol. Thus, the regulation by nucleotides does not appear to reflect binding at a separate regulatory site. The random kinetic mechanism obviates the need to postulate such a site to explain the formation of binary complexes with the nucleotides. The observed stoichiometry is consistent with a model for the nucleotide regulatory behavior in which the dimer is the enzyme form present in the assay and its subunits display different substrate binding affinities. Several properties of the enzyme are consistent with negative cooperativity as the basis for the difference in affinities. The possible physiological importance of the regulatory behavior with respect to ATP is considered.  相似文献   

20.
Yu P  Lasagna M  Pawlyk AC  Reinhart GD  Pettigrew DW 《Biochemistry》2007,46(43):12355-12365
Steady-state and time-resolved fluorescence anisotropy methods applied to an extrinsic fluorophore that is conjugated to non-native cysteine residues demonstrate that amino acids in an allosteric communication network within a protein subunit tune protein backbone motions at a distal site to enable allosteric binding and inhibition. The unphosphorylated form of the phosphocarrier protein IIAGlc is an allosteric inhibitor of Escherichia coli glycerol kinase, binding more than 25 A from the kinase active site. Crystal structures that showed a ligand-dependent conformational change and large temperature factors for the IIAGlc-binding site on E. coli glycerol kinase suggest that motions of the allosteric site have an important role in the inhibition. Three E. coli glycerol kinase amino acids that are located at least 15 A from the active site and the allosteric site were shown previously to be necessary for transplanting IIAGlc inhibition into the nonallosteric glycerol kinase from Haemophilus influenzae. These three amino acids are termed the coupling locus. The apparent allosteric site motions and the requirement for the distant coupling locus to transplant allosteric inhibition suggest that the coupling locus modulates the motions of the IIAGlc-binding site. To evaluate this possibility, variants of E. coli glycerol kinase and the chimeric, allosteric H. influenzae glycerol kinase were constructed with a non-native cysteine residue replacing one of the native residues in the IIAGlc-binding site. The extrinsic fluorophore Oregon Green 488 (2',7'-difluorofluorescein) was conjugated specifically to the non-native cysteine residue. Steady-state and time-resolved fluorescence anisotropy measurements show that the motions of the fluorophore reflect backbone motions of the IIAGlc-binding site and these motions are modulated by the amino acids at the coupling locus.  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