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.
Studies of the interaction between actin and myosin subfragment 1 (S1) in solution have shown that the association reaction takes place in at least two steps. Initially the association is relatively weak to form a complex called the A state which can then isomerize to the R state. The rate and equilibrium constants for the isomerization have been measured and are shown to depend upon the nucleotide bound to the S1 ATPase site; with ATP bound the A state is preferred but as ATP is hydrolysed and the products are sequentially released then the complex gradually shifts to the A state. An extensive series of experiments have characterized the A-to-R isomerization both in solution and in contracting muscle fibres and have shown it to be closely associated with the key events in the ATP-driven contraction cycle: the conformational change from the A to the R state can be monitored by fluorescent probes on either actin or the nucleotide; the isomerization can be perturbed by increases in hydrostatic pressure; the actin-induced acceleration of the rate of product release from myosin is coupled to the A-to-R isomerization; tropomyosin may control actin and myosin interaction by controlling the isomerization step and finally pressure perturbations of contracting muscle fibres shows there to be a close coupling between the isomerization of acto.S1 and the force generating event of muscle contraction.  相似文献   

2.
We have examined the kinetics of nucleotide binding to actomyosin VI by monitoring the fluorescence of pyrene-labeled actin filaments. ATP binds single-headed myosin VI following a two-step reaction mechanism with formation of a low affinity collision complex (1/K(1)' = 5.6 mm) followed by isomerization (k(+2)' = 176 s-1) to a state with weak actin affinity. The rates and affinity for ADP binding were measured by kinetic competition with ATP. This approach allows a broader range of ADP concentrations to be examined than with fluorescent nucleotide analogs, permitting the identification and characterization of transiently populated intermediates in the pathway. ADP binding to actomyosin VI, as with ATP binding, occurs via a two-step mechanism. The association rate constant for ADP binding is approximately five times greater than for ATP binding because of a higher affinity in the collision complex (1/K(5b)' = 2.2 mm) and faster isomerization rate constant (k(+5a)' = 366 s(-1)). By equilibrium titration, both heads of a myosin VI dimer bind actin strongly in rigor and with bound ADP. In the presence of ATP, conditions that favor processive stepping, myosin VI does not dwell with both heads strongly bound to actin, indicating that the second head inhibits strong binding of the lead head to actin. With both heads bound strongly, ATP binding is accelerated 2.5-fold, and ADP binding is accelerated >10-fold without affecting the rate of ADP release. We conclude that the heads of myosin VI communicate allosterically and accelerate nucleotide binding, but not dissociation, when both are bound strongly to actin.  相似文献   

3.
V N Kasho  M Yoshida  P D Boyer 《Biochemistry》1989,28(17):6949-6954
The ATPase from the ATP synthase of the thermophilic bacterium PS3 (TF1), unlike F1 ATPase from other sources, does not retain bound ATP, ADP, and Pi at a catalytic site under conditions for single-site catalysis [Yohda, M., & Yoshida, M. (1987) J. Biochem. 102, 875-883]. This raised a question as to whether catalysis by TF1 involved alternating participation of catalytic sites. The possibility remained, however, that there might be transient but catalytically significant retention of bound reactants at catalytic sites when the medium ATP concentration was relatively low. To test for this, the extent of water oxygen incorporation into Pi formed by ATP hydrolysis was measured at various ATP concentrations. During ATP hydrolysis at both 45 and 60 degrees C, the extent of water oxygen incorporation into the Pi formed increased markedly as the ATP concentration was lowered to the micromolar range, with greater modulation observed at 60 degrees C. Most of the product Pi formed arose by a single catalytic pathway, but measurable amounts of Pi were formed by a pathway with high oxygen exchange. This may result from the presence of some poorly active enzyme. The results are consistent with sequential participation of three catalytic sites on the TF1 as predicted by the binding change mechanism.  相似文献   

4.
The oxygen exchange parameters for the hydrolysis of ATP by the F1-ATPase have been determined over a 140,000-fold range of ATP concentrations and a 5,000-fold range of reaction velocity. The average number of water oxygens incorporated into each Pi product ranges from a limit of about 1.02 at saturating ATP concentrations to a limit of about 3.97 at very low ATP concentrations. The latter value represents 400 reversals of hydrolysis of bound ATP prior to Pi dissociation. In accord with the binding change mechanism, this means that ATP binding at one catalytic site increases the off constant of Pi and ADP from another catalytic site by at least 20,000-fold, equivalent to the use of 6 kcal mol-1 of ATP binding energy to promote product release. The estimated rate of reversal of hydrolysis of F1-ATPase-bound ATP to bound ADP + Pi varies only about 5-fold with ATP concentration. The rate is similar that observed previously for reversal of bound ATP hydrolysis or synthesis with the membrane-bound enzyme and is greater than the rate of net ATP formation during oxidative phosphorylation. This adds to evidence that energy input or membrane components are not required for bound ATP synthesis.  相似文献   

5.
I A Rose 《Federation proceedings》1978,37(14):2775-2782
Reversible gamma-PO3 transfer in ATP reactions can be recognized by exchange of 18O from the beta,gamma-bridge position to the beta-P-nonbridge positions: (see article). Such intramolecular exchange is less demanding for the detection of the bond cleavage than the usual ATP:ADP isotope exchange because it does not require dissociation of bound ADP from the intermediate complex. Acyl phosphate intermediates are indicated for the glutamine synthetase and carbamyl-P synthetase reactions by their extreme requirements for glutamate and bicarbonate, respectively, for positional oxygen exchange. No support is given for E-P or concerted mechanisms. No support is found for an active CO2 in the latter reaction, although this is not ruled out by the data. Positional isomerization in ATP occurs with lamellae from spinach chloroplast only in the light. When the ATP molecule interacts, it also undergoes complete exchange of the gamma-PO3 oxygen with water before it rejoins the pool of free ATP. The difference in rates of the two exchanges suggests that the torsional motion of ADP-beta-PO3 is greatly hindered on the enzyme. This may explain, by the argument of substrate activation, the rapid reversibility of the ATPase reaction on the enzyme.  相似文献   

6.
Raman spectroscopy has been used to determine the isotopic distribution of oxygen during the isomerization of pernitrous acid to nitrate in18OH2. Decomposition of unlabeled pernitrous acid in carbon-dioxide-free phosphate-buffered solutions of18OH2at pH 6.8 results in the incorporation of 13 ± 1%18O into the nitrate product with no detectable double-isotopic incorporation and 83% of the product having complete retention of its oxygen atoms. The mechanistic implications of this are contrasted for three frequently considered pathways for peroxide bond cleavage and it is concluded that the results are most consistent with the formation of reactive intermediate which can either rearrange or undergo hydrolytic O-atom exchange with water.  相似文献   

7.
Evidence is presented that the myosin subfragment-1–ADP complex, generated by the addition of Mg2+ and ADP to subfragment 1, is an intermediate within the myosin Mg2+-dependent adenosine triphosphatase (ATPase) turnover cycle. The existence of this species as a steady-state intermediate at pH8 and 5°C is demonstrated by fluorescence measurements, but its concentration becomes too low to measure at 21°C. This arises because there is a marked temperature-dependence on the rate of the process controlling ADP dissociation from subfragment 1 (rate=1.4s−1 at 21°C, 0.07s−1 at 5°C). In the ATPase pathway this reaction is in series with a relatively temperature-insensitive process, namely an isomerization of the subfragment-1–product complex (rate=0.055s−1 at 21°C, 0.036s−1 at 5°C). By means of studies on the Pi inhibition of nucleotide-association rates, a myosin subfragment-1–Pi complex was characterized with a dissociation equilibrium constant of 1.5mm. Pi appears to bind more weakly to the myosin subfragment-1–ADP complex. The studies indicate that Pi dissociates from subfragment 1 at a rate greater than 40s−1, and substantiates the existence of a myosin-product isomerization before product release in the elementary processes of the Mg2+-dependent ATPase. In this ATPase mechanism Mg2+ associates as a complex with ATP and is released as a complex with ADP. In 0.1m-KCl at pH8 1.0mol of H+ is released/mol of subfragment 1 concomitant with the myosin-product isomerization or Pi dissociation, and 0.23 mol of H+ is released/mol of subfragment when ATP binds to the protein, but 0.23 mol of H+ is taken up again from the medium when ADP dissociates. Within experimental sensitivity no H+ is released into the medium in the step involving ATP cleavage.  相似文献   

8.
Cochran JC  Gilbert SP 《Biochemistry》2005,44(50):16633-16648
The ATPase mechanism of kinesin superfamily members in the absence of microtubules remains largely uncharacterized. We have adopted a strategy to purify monomeric human Eg5 (HsKSP/Kinesin-5) in the nucleotide-free state (apoEg5) in order to perform a detailed transient state kinetic analysis. We have used steady-state and presteady-state kinetics to define the minimal ATPase mechanism for apoEg5 in the absence and presence of the Eg5-specific inhibitor, monastrol. ATP and ADP binding both occur via a two-step process with the isomerization of the collision complex limiting each forward reaction. ATP hydrolysis and phosphate product release are rapid steps in the mechanism, and the observed rate of these steps is limited by the relatively slow isomerization of the Eg5-ATP collision complex. A conformational change coupled to ADP release is the rate-limiting step in the pathway. We propose that the microtubule amplifies and accelerates the structural transitions needed to form the ATP hydrolysis competent state and for rapid ADP release, thus stimulating ATP turnover and increasing enzymatic efficiency. Monastrol appears to bind weakly to the Eg5-ATP collision complex, but after tight ATP binding, the affinity for monastrol increases, thus inhibiting the conformational change required for ADP product release. Taken together, we hypothesize that loop L5 of Eg5 undergoes an "open" to "closed" structural transition that correlates with the rearrangements of the switch-1 and switch-2 regions at the active site during the ATPase cycle.  相似文献   

9.
Oxidative stress can damage the active site cysteine of the antioxidant enzyme peroxiredoxin (Prx) to the sulfinic acid form, Prx-SO2. This modification leads to inactivation. Sulfiredoxin (Srx) utilizes a unique ATP-Mg2+-dependent mechanism to repair the Prx molecule. Using selective protein engineering that involves disulfide bond formation and site-directed mutagenesis, a mimic of the enzyme·substrate complex has been trapped. Here, we present the 2.1 Å crystal structure of human Srx in complex with PrxI, ATP, and Mg2+. The Cys52 sulfinic acid moiety was substituted by mutating this residue to Asp, leading to a replacement of the sulfur atom with a carbon atom. Because the Srx reaction cannot occur, the structural changes in the Prx active site that lead to the attack on ATP may be visualized. The local unfolding of the helix containing C52D resulted in the packing of Phe50 in PrxI within a hydrophobic pocket of Srx. Importantly, this structural rearrangement positioned one of the oxygen atoms of Asp52 within 4.3 Å of the γ-phosphate of ATP bound to Srx. These observations support a mechanism where phosphorylation of Prx-SO2 is the first chemical step.  相似文献   

10.
The adenosine 5'-O-(3-thiotriphosphate) (ATP gamma S) induced dissociation of actomyosin subfragment 1 (S1) has been investigated by monitoring the light scattering changes that occur on dissociation. We have shown that ATP gamma S dissociates acto-S1 by a mechanism similar to that of ATP but at a rate 10 times slower. The maximum rate of dissociation is limited by an isomerization of the ternary actin-S1-nucleotide complex, which has a rate of 500 s-1 for ATP gamma S and an estimated rate of 5000 s-1 for ATP (20 degrees C, 0.1 M KCl, pH 7.0). The activation energy for the isomerization is the same for ATP and ATP gamma S, and both show a break in the Arrhenius plot at 5 degrees C. The reaction between acto-S1 and ATP was also followed by the fluorescence of a pyrene group covalently attached to Cys-374. We show that the fluorescence of the pyrene group reports the isomerization step and not actin dissociation. The characterization of this isomerization is discussed in relation to force-generating models of the actomyosin cross-bridge cycle.  相似文献   

11.
The forward and reverse rates of the overall reaction catalyzed by the ATP synthase in intact rat heart mitochondria, as measured with 32P, were compared with the rates of two partial steps, as measured with 18O. Such rates have been measured previously, but their relationship to one another has not been determined, nor have the partial reactions been measured in intact mitochondria. The partial steps measured were the rate of medium Pi formation from bound ATP (in state 4 this also equals the rate of medium Pi into bound ATP) and the rate of formation of bound ATP from bound Pi within the catalytic site. The rates of both partial reactions can be measured by 31P NMR analysis of the 18O distribution in Pi and ATP released from the enzyme during incubation of intact mitochondria with highly labeled [18O]Pi. Data were obtained in state 3 and 4 conditions with variation in substrate concentrations, temperature, and mitochondrial membrane electrical potential gradient (delta psi m). Although neither binding nor release of ATP is necessary for phosphate/H2O exchange, in state 4 the rate of incorporation of at least one water oxygen atom into phosphate is approximately twice the rate of the overall reaction rate under a variety of conditions. This can be explained if the release of Pi or ATP at one catalytic site does not occur, unless ATP or Pi is bound at another catalytic site. Such coupling provides strong support for the previously proposed alternating site mechanism. In state 3 slow reversal of ATP synthesis occurs within the mitochondrial matrix and can be detected as incorporation of water oxygen atoms into medium Pi even though medium [32P]ATP does not give rise to 32Pi in state 3. These data can be explained by lack of translocation of ATP from the medium to the mitochondrial matrix. The rate of bound ATP formation from bound Pi at catalytic sites was over twice the rate of the overall reaction in both states 4 and 3. The rate of reaction at the catalytic site is considerably less sensitive to the decrease in membrane potential and the concentration of medium ADP than is the rate of medium ATP formation. This supports the view that the active catalytic site is occluded and proceeds at a rapid rate which is relatively independent of delta psi m and of media substrates.  相似文献   

12.
Two interesting previously reported properties of mitochondrial F1 ATPase have been confirmed and have been examined by18O exchange measurements to assess if they are consistent with sequential participation of catalytic sites during ATP hydrolysis. These are the ability of HCO 3 to increase reaction rate with apparent loss of cooperative interaction between subunits and the ability of ITP to accelerate the hydrolysis of a low concentration of ATP. The effect of HCO 3 was tested at concentrations of ATP lower than previous measurements. The activation disappeared when ATP was reduced to 0.1 µM. The HCO 3 activation at higher ATP concentrations did not change the extent of reversal of the cleavage of tightly bound ATP at the catalytic site, as measured by the average number of water oxygens incorporated with each Pi formed when 5 or 10 µM ATP is hydrolyzed. The data are consistent with sequential site participation with HCO 3 acceleration of ADP departure after a binding change that stops18O exchange and loosens ADP binding.When ITP concentration was lowered during net ITP hydrolysis by F1 ATPase an increase in water oxygen incorporation into Pi formed is observed, as noted previously for ATP hydrolysis. The acceleration of the cleavage of a constant low concentration of [-18O]ATP by concomitant hydrolysis of increasing concentrations of ITP was accompanied by a decrease in water oxygen incorporation with each Pi formed from the ATP. These results add to evidence for the binding change mechanism for F1 ATPase with sequential participation of catalytic sites.  相似文献   

13.
Gadda G  Fitzpatrick PF 《Biochemistry》2000,39(6):1400-1405
The flavoprotein nitroalkane oxidase catalyzes the oxidation of primary and secondary nitroalkanes to aldehydes and ketones, respectively, transferring electrons to oxygen to form hydrogen peroxide. The steady-state kinetic mechanism of the active flavin adenine dinucleotide-(FAD-) containing form of the enzyme has been determined with nitroethane at pH 7 to be bi-ter ping-pong, with oxygen reacting with the free reduced enzyme after release of the aldehyde product. The V(max) value is 5.5 +/- 0.3 s(-)(1) and the K(m) values for nitroethane and oxygen are 3.3 +/- 0.6 and 0.023 +/- 0.007 mM, respectively. The free reduced enzyme forms a dead-end complex with nitroethane, with a K(ai) value of 30 +/- 6 mM. Acetaldehyde and butyraldehyde are noncompetitive inhibitors versus nitroethane due to formation of a dead-end complex between the oxidized enzyme and the product. Acetaldehyde is an uncompetitive inhibitor versus oxygen, indicating that an irreversible isomerization of the free reduced enzyme occurs before the reaction with oxygen. Addition of unprotonated imidazole results in a 5-fold increase in the V(max) value, while the V/K values for nitroethane and oxygen are unaffected. A 5-fold increase in the K(ai) value for nitroethane and a 6.5-fold increase in the K(ii) value for butyraldehyde are observed in the presence of imidazole. These results are consistent with the isomerization of the free reduced enzyme being about 80% rate-limiting for catalysis and with a model in which unprotonated imidazole accelerates the rate of isomerization.  相似文献   

14.
To study the catalytic mechanism of phosphorylation catalyzed by cAMP-dependent protein kinase (PKA) a structure of the enzyme-substrate complex representing the Michaelis complex is of specific interest as it can shed light on the structure of the transition state. However, all previous crystal structures of the Michaelis complex mimics of the PKA catalytic subunit (PKAc) were obtained with either peptide inhibitors or ATP analogs. Here we utilized Ca2+ ions and sulfur in place of the nucleophilic oxygen in a 20-residue pseudo-substrate peptide (CP20) and ATP to produce a close mimic of the Michaelis complex. In the ternary reactant complex, the thiol group of Cys-21 of the peptide is facing Asp-166 and the sulfur atom is positioned for an in-line phosphoryl transfer. Replacement of Ca2+ cations with Mg2+ ions resulted in a complex with trapped products of ATP hydrolysis: phosphate ion and ADP. The present structural results in combination with the previously reported structures of the transition state mimic and phosphorylated product complexes complete the snapshots of the phosphoryl transfer reaction by PKAc, providing us with the most thorough picture of the catalytic mechanism to date.  相似文献   

15.
The cyanobacterial circadian clock oscillator is composed of three clock proteins—KaiA, KaiB, and KaiC, and interactions among the three Kai proteins generate clock oscillation in vitro. However, the regulation of these interactions remains to be solved. Here, we demonstrated that ATP regulates formation of the KaiB-KaiC complex. In the absence of ATP, KaiC was monomeric (KaiC1mer) and formed a complex with KaiB. The addition of ATP plus Mg2+ (Mg-ATP), but not that of ATP only, to the KaiB-KaiC1mer complex induced the hexamerization of KaiC and the concomitant release of KaiB from the KaiB-KaiC1mer complex, indicating that Mg-ATP and KaiB compete each other for KaiC. In the presence of ATP and Mg2+ (Mg-ATP), KaiC became a homohexameric ATPase (KaiC6mer) with bound Mg-ATP and formed a complex with KaiB, but KaiC hexamerized by unhydrolyzable substrates such as ATP and Mg-ATP analogs, did not. A KaiC N-terminal domain protein, but not its C-terminal one, formed a complex with KaiB, indicating that KaiC associates with KaiB via its N-terminal domain. A mutant KaiC6mer lacking N-terminal ATPase activity did not form a complex with KaiB whereas a mutant lacking C-terminal ATPase activity did. Thus, the N-terminal domain of KaiC is responsible for formation of the KaiB-KaiC complex, and the hydrolysis of the ATP bound to N-terminal ATPase motifs on KaiC6mer is required for formation of the KaiB-KaiC6mer complex. KaiC6mer that had been hexamerized with ADP plus aluminum fluoride, which are considered to mimic ADP-Pi state, formed a complex with KaiB, suggesting that KaiB is able to associate with KaiC6mer with bound ADP-Pi.  相似文献   

16.
A new approach to the direct estimation of the value of the off constant for dissociation of ATP from myosin subfragment 1 (S1) has been developed. From measurements of the extremely slow rate of release of [32P]-ATP formed from 32Pi by S1 catalysis and the amount of rapidly formed [32P]-ATP tightly bound to S1, the value of the off constant is approximately 2.8 × 10−4 sec−1 at pH 7.4. The concentration dependencies for Pi ⇌ H18 OH exchange and for 32Pi incorporation into myosin-bound ATP give direct measurements of the dissociation constant of Pi from S1. Both approaches show that the enzyme has a very low affinity for Pi, with an apparent Kd of > 400 mM. Measurement of the average number of water oxygens incorporated into Pi released from ATP by S1-catalyzed hydrolysis in the presence of Mg2+ suggests that the hydrolytic step reverses an average of at least 5.5 times for each ATP cleaved. With the Ca2+-activated hydrolysis, less than one oxygen from water appears in each Pi released. This finding is indicative of a possible isotope effect in the attack of water on the terminal phosphoryl group of ATP.  相似文献   

17.
The initial rates of ATP hydrolysis catalyzed by Fo x F1 (bovine heart submitochondrial particles) preincubated in the presence of Pi for complete activation of the oligomycin-sensitive ATPase were measured as a function of ATP, Mg2+, and Mg x ATP concentrations. The results suggest the mechanism in which Mg x ATP complex is the true substrate of the ATPase and the second Mg2+ bound at a specific pH-dependent site is needed for the catalysis. Simple hyperbolic Michaelis--Menten dependences of the reaction rate on the substrate (Mg x ATP) and activating Mg2+ were found. In contrast to the generally accepted view, no inhibition of ATPase by free Mg2+ was found. Inhibition of the reaction by free ATP is due to a decrease of free Mg2+ needed for the catalysis. In the presence of both Ca2+ and Mg2+ the kinetics of ATP hydrolysis suggest that the Ca x ATP complex is neither hydrolyzed nor competes with Mg x ATP, and free Ca2+ does not affect the hydrolysis of Mg x ATP complex. A crucial role of free Mg2+ in the time-dependent inhibition of ATPase by azide is shown. The dependence of apparent Km for Mg x ATP on saturation of the Mg2+-specific site suggests the formal ping-pong mechanism in which bound Mg2+ participates in the overall reaction after dissociation of one product (most likely Pi) thus promoting either release of ADP (catalytic turnover) or slow isomerization of the enzyme--product complex (formation of the dead-end ADP(Mg2+)-inhibited enzyme). The rate of Mg x ATP hydrolysis only slightly depends on pH at saturating Mg2+. In the presence of limited amounts of free Mg2+ the pH dependence of the initial rate corresponds to the titration of a single group with pKa = 7.5. The simple competition between H+ and activating Mg2+ was observed. The specific role of Mg2+ as a coupling cation for energy transduction in Fo x F1-ATPase is discussed.  相似文献   

18.
Binding of 1,N6-ethanoadenosine triphosphate to actin   总被引:3,自引:0,他引:3  
G-actin is known to bind one molecule of ATP. Its polymerization to F-actin is accompanied by the splitting off of the terminal phosphate of the bound nucleotide. We have found that the fluorescent 1,N6-ethanoadenosine triphosphate (?ATP) can substitute for ATP in G-actin and that G-actin containing bound ?ATP possesses essentially full polymerizability. The binding of this ATP analog has been studied by following the inactivation of the ?ATP·G-actin complex. The binding constant (4?5.7 × 106 M?1) obtained in the absence of EDTA is about 50% of that for ATP, while the binding constant obtained in the presence of EDTA (0.9?3.0 × 105 M?1) is comparable to those for ATP and ADP. These findings suggest that ?ATP can be used as a structural probe for actin. The fluorescence lifetime of ?ATP bound to G·actin is 36 nsec. The rotational relaxation time of ?ATP·G-actin is near 60 nsec. at 20°C.  相似文献   

19.
Kim J  Tsai PC  Chen SL  Himo F  Almo SC  Raushel FM 《Biochemistry》2008,47(36):9497-9504
The bacterial phosphotriesterase (PTE) from Pseudomonas diminuta catalyzes the hydrolysis of organophosphate esters at rates close to the diffusion limit. X-ray diffraction studies have shown that a binuclear metal center is positioned in the active site of PTE and that this complex is responsible for the activation of the nucleophilic water from solvent. In this paper, the three-dimensional structure of PTE was determined in the presence of the hydrolysis product, diethyl phosphate (DEP), and a product analogue, cacodylate. In the structure of the PTE-diethyl phosphate complex, the DEP product is found symmetrically bridging the two divalent cations. The DEP displaces the hydroxide from solvent that normally bridges the two divalent cations in structures determined in the presence or absence of substrate analogues. One of the phosphoryl oxygen atoms in the PTE-DEP complex is 2.0 A from the alpha-metal ion, while the other oxygen is 2.2 A from the beta-metal ion. The two metal ions are separated by a distance of 4.0 A. A similar structure is observed in the presence of cacodylate. Analogous complexes have previously been observed for the product complexes of isoaspartyl dipeptidase, d-aminoacylase, and dihydroorotase from the amidohydrolase superfamily of enzymes. The experimentally determined structure of the PTE-diethyl phosphate product complex is inconsistent with a recent proposal based upon quantum mechanical/molecular mechanical simulations which postulated the formation of an asymmetrical product complex bound exclusively to the beta-metal ion with a metal-metal separation of 5.3 A. This structure is also inconsistent with a chemical mechanism for substrate hydrolysis that utilizes the bridging hydroxide as a base to abstract a proton from a water molecule loosely associated with the alpha-metal ion. Density functional theory (DFT) calculations support a reaction mechanism that utilizes the bridging hydroxide as the direct nucleophile in the hydrolysis of organophosphate esters by PTE.  相似文献   

20.
Under steady state photophosphorylating conditions, each ATP synthase complex from spinach thylakoids contains, at a catalytic site, about one tightly bound ATP molecule that is rapidly labeled from medium 32Pi. The level of this bound [32P]ATP is markedly reduced upon de-energization of the spinach thylakoids. The reduction is biphasic, a rapid phase in which the [32P] ATP/synthase complex drops about 2-fold within 10 s, followed by a slow phase, kobs = 0.01/min. A decrease in the concentration of medium 32Pi to well below its apparent Km for photophosphorylation is required to decrease the amount of tightly bound ATP/synthase found just after de-energization and before the rapid phase of bound ATP disappearance. The [32P]ATP that remains bound after the rapid phase appears to be mostly at a catalytic site as demonstrated by a continued exchange of the oxygens of the bound ATP with water oxygens. This bound [32P]ATP does not exchange with medium Pi and is not removed by the presence of unlabeled ATP. The levels of tightly bound ADP and ATP arising from medium ADP were measured by a novel method based on use of [beta-32P]ADP. After photophosphorylation and within minutes after the rapid phase of bound ATP loss, the measured ratio of bound ADP to ATP was about 1.4 and the sum of bound ADP plus ATP was about 1/synthase. This ratio is smaller than that found about 1 h after de-energization. Hence, while ATP bound at catalytic sites disappears, bound ADP appears. The results suggest that during and after de-energization the bound ATP disappears from the catalytic site by hydrolysis to bound ADP and Pi with subsequent preferential release of Pi. These and related observations can be accommodated by the binding change mechanism for ATP synthase with participation of alternating catalytic sites and are consistent with a deactivated state arising from occupancy of one catalytic site on the synthase complex by an inhibitory ADP without presence of Pi.  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